Next Article in Journal
The Role of Diacylglycerol Kinase in the Amelioration of Diabetic Nephropathy
Next Article in Special Issue
Formation Optimization, Characterization and Antioxidant Activity of Auricularia auricula-judae Polysaccharide Nanoparticles Obtained via Antisolvent Precipitation
Previous Article in Journal
G-Quadruplex Aptamer-Ligand Characterization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

MoS2 and MoS2 Nanocomposites for Adsorption and Photodegradation of Water Pollutants: A Review

by
Leonardo O. Amaral
and
Ana L. Daniel-da-Silva
*
Department of Chemistry and CICECO-Aveiro Institute of Materials, University of Aveiro, 3810-193 Aveiro, Portugal
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(20), 6782; https://doi.org/10.3390/molecules27206782
Submission received: 20 September 2022 / Revised: 4 October 2022 / Accepted: 7 October 2022 / Published: 11 October 2022
(This article belongs to the Special Issue Women in Nanochemistry)

Abstract

:
The need for fresh and conveniently treated water has become a major concern in recent years. Molybdenum disulfide (MoS2) nanomaterials are attracting attention in various fields, such as energy, hydrogen production, and water decontamination. This review provides an overview of the recent developments in MoS2-based nanomaterials for water treatment via adsorption and photodegradation. Primary attention is given to the structure, properties, and major methods for the synthesis and modification of MoS2, aiming for efficient water-contaminant removal. The combination of MoS2 with other components results in nanocomposites that can be separated easily or that present enhanced adsorptive and photocatalytic properties. The performance of these materials in the adsorption of heavy metal ions and organic contaminants, such as dyes and drugs, is reviewed. The review also summarizes current progress in the photocatalytic degradation of various water pollutants, using MoS2-based nanomaterials under UV-VIS light irradiation. MoS2-based materials showed good activity after several reuse cycles and in real water scenarios. Regarding the ecotoxicity of the MoS2, the number of studies is still limited, and more work is needed to effectively evaluate the risks of using this nanomaterial in water treatment.

1. Introduction

The world’s population has been growing yearly, and an increased need for resources has accompanied this growth [1]. Industrial, agricultural, and medical development have evolved to supply this population growth [1,2]. However, after this substantial increase in resource consumption, pollution had also increased, becoming a huge concern, along with the need for freshwater sources or for effective water treatment [1,3]. In this regard, several approaches for water treatment have been developed, such as adsorption and photocatalysis, to remove pollutants and other threats to human health and ecosystems, such as heavy metals, dyes, drugs, and pesticides, among others [4,5,6,7,8].
Along with the advances in environmental research, in the past few years, several authors have suggested nanomaterials to enhance primary and advanced water treatment [9]. Due to its small size and, consequently, high surface and reactive areas, the use of nanomaterials enables the improvement of several properties compared to bulk material [9]. More recently, two-dimensional (2D) nanomaterials have gained importance. In this group are included materials such as graphene, hexagonal boron nitride, 2D honeycomb silicon, and transition metal dichalcogenides (TMD) such as molybdenum disulfide (MoS2) and tungsten disulfide (WS2) [10]. TMD are hexagonal, structured and layered materials with the molecular formula MX2, where M is a transition metal (Mo, W, or Nb) and X represents a chalcogen element (S, Se, or Te) [11,12]. Within this group of materials, nano-MoS2 has gained attention year after year due to its outstanding chemical, electronic, catalytic, optical, and mechanical properties, making it suitable for various fields such as pollution remediation and sensors, medicine, mechanics and electronics, and energy conversion and storage [13,14,15,16,17,18,19,20,21,22,23,24]. When compared to other sulfides, such as zinc sulfide (ZnS), cadmium sulfide (CdS), and tungsten sulfide (WS2), MoS2 show promising photocatalytic activity, allied to its easy preparation and unique optical and electronic properties [25]. The wide band gaps of ZnS and MnS render them responsive to only UV light; further modification is required to extend the light absorption range [26]. In contrast, MoS2 is active with visible light. As a photocatalytic material, MoS2 presents a suitable surface area with a large number of active site edges, a layered structure that allows good adsorption capacity, and a tunable band gap between 1.2 eV and 1.9 eV, depending on the number of layers [27]. Regarding safety considerations, Mo is less hazardous and can be used to substitute Cd, which is known to be a human carcinogen.
In this review, we first summarize the main structures and properties of MoS2 and the most common methods by which to prepare MoS2-based materials. We then critically review the recent advances in MoS2-based materials for water treatment and purification through adsorption and photocatalysis.

2. MoS2 Structure and Properties

In nature, MoS2 exists in molybdenite bulk mineral form as a black powder [28]. Bulk MoS2 can be exfoliated, obtaining few-layer nanosheets (FLMoS2) and single-layer nanosheets (SLMoS2) [29,30]. SLMoS2 compared to bulk MoS2 has intrinsic higher electronic conductivity and presents a direct gap of 1.7–1.9 eV, while bulk MoS2 presents an indirect band gap of 1.1–1.3 eV [31,32,33,34]. This relatively small band gap of SLMoS2 enables the use of visible light to perform photocatalytic reactions [35,36]. While absent in the bulk material, SLMoS2 also shows strong photoluminescence [37]. At room temperature, SLMoS2 exhibits an on/off current ratio of 108 and an electron mobility of 200 cm2 (V.s)−1 [38]. Notably, several properties of MoS2 can be adjusted by varying the number of layers and polymorphism. SLMoS2 presents a sandwich structure with a thickness of 0.6 nm, composed of two layers of atomic S, with a single layer of atomic Mo between, linked through Mo-S covalent bonds [39,40,41]. In the case of FLMoS2, the different sheets interact with each other via weak van der Waals (vdW) forces [40,42]. Because of these weak vdW forces, changing SLMoS2 to FLMoS2 and gaining the reverse reaction is easily achieved [41]. Kumar and Mishra demonstrated, via molecular simulation, the mechanism of solvent-assisted exfoliation to turn bulk MoS2 into FLMoS2 by applying a shock-wave method [43].
MoS2 presents several polymorphs, depending on the interlayer stacking arrangement and intralayer coordination between the central Mo atom and the surrounding S atoms, the crystal structures of which are shown in Figure 1 [41,44,45]. The MoS2 polymorphs are identified using an adapted version of Ramsdell’s notation that was introduced for SiC polytypes [46]. This nomenclature indicates the number of S-Mo-S layers in the unit cell (1, 2, or 3) and the crystal system of the lattice, where “T”, “H”, and “R” stand for trigonal, hexagonal, and rhombohedral, respectively [47]. Five polymorphic forms have been identified for MoS2: 1T, 1T’, 1H, 2H, and 3R [4,33,44,48,49,50,51,52,53,54,55]. The forms 1T, 2H, and 3R are polytypes because they differ in terms of stacking configuration. The polytypes 2H and 3R occur naturally, while 1T is a metastable structure. The semiconducting 2H phase is naturally more stable and has been extensively investigated for energy-related applications [56,57], transistors [58], photodetectors [59], valleytronics, spintronic devices, and optoelectric devices [58,60,61]. In the 2H-MoS2 phase, each Mo coordinates with S atoms in a trigonal prismatic arrangement, with two layers in each unit cell in hexagonal symmetry, as shown in Figure 1. The semiconducting behavior of 2H-MoS2 results from a finite band gap between filled dz2 and empty dx2−y2,xy bands [62]. The 2H-MoS2 phase can easily be turned into the 1T phase by exfoliating the 2H-MoS2 nanosheets, using the sonication-assisted lithium intercalation method, and treating the resulting material with an infrared laser [63].
With the same prismatic coordination as the 2H MoS2, but with three stacking layers of S–Mo–S directed along the c-axis instead of two (Figure 1), the rhombohedral 3R phase is also a naturally occurring semiconducting phase. Due to the broken inversion symmetry, this phase has drawn extensive attention because of its potential applications for non-linear optical devices [64,65,66]. Compared to monolayers, the 3R phase presents a higher predicted piezoelectricity coefficient [64]. The 3R-phase can be synthesized via homoepitaxial growth through chemical vapor deposition [67,68]. Through heating, the 3R phase can easily turn into the 2H phase [69].
The monolayered 1T, 1H and 1T’-MoS2 phases are other polytypes that have attracted attention but do not occur naturally. Depending on the S-atom’s geometry coordination, the Mo atom can be coordinated by six S atoms in either an octahedral (1T) or trigonal prismatic (1H) arrangement [44]. The 1T-phase can be obtained via exfoliation of the 2H-phase [70]. It is metallic in nature, with an electrical conductivity 107 times higher than the 2H phase, and is of interest as a supercapacitor electrode material [70,71]. This metastable and paramagnetic phase has been reported to be efficient as a photocatalyst for hydrogen evolution and gas molecule adsorption, due to the high surface activity at the basal sites [44,45,52,53,62,70]. It is also known that the 1T phase can turn into the 2H phase with high-temperature treatment [62]. The semimetallic 1T’-phase, the more stable SLMoS2 with a distorted structure, due to the clusterization of Mo atoms, is formed through the dimerization of Mo atoms in the 1T phase [54]. The 1H phase is a semiconductor. It is possible to turn 1H-MoS2 into 1T-MoS2 by applying the Li intercalation method, in which an electron is transferred from the alkali metal to the d orbital of the transition metal center, resulting in the metallic-like character of the material [54].
The phase composition influences the photocatalytic performance of MoS2 nanosheets. Materials comprising 1T/2H phases have shown higher performance in the photodegradation of methyl orange dye, compared to the 2H and 3R phases [69]. Computational studies have indicated that this enhanced photocatalytic activity is due to the emergence of mid-gap states upon the introduction of 1T sites to the 2H lattice. Density functional theory (DFT)-based calculations investigating the material’s structural and electronic properties found a zero tunnel-barrier height in stable 2H/1T heterostructures, which is advantageous for achieving an efficient carrier injection rate [72].
Regarding the adsorptive properties, according to the principle of a “hard and soft acid and basis” (HSAB), soft acids (such as heavy metal ions) are capable of strong interaction with soft bases, such as sulfur atoms [73]. Thus, the MoS2 nanosheets show high adsorption capacity due to the abundance of adsorption sites and the fast kinetics caused by easy access to those sites [74]. Concerning solubility, MoS2 is soluble in aqua regia and in hot concentrated sulfuric acid, but it shows insolubility in water and diluted acid, indicating long-term stability in water systems [75].

3. Common Synthesis Methods and Modifications of MoS2

MoS2 nanosheets can be synthesized through top-down methods, which consist of exfoliating the bulk MoS2, and the bottom-up approach, consisting of growing nanosheets from single atoms using chemical methods such as chemical vapor deposition (CVD) or the hydrothermal method. Exfoliation processes are highly efficient and conveniently scalable, producing by-products that are less hazardous and lower costs than bottom-up methods [76,77]. However, bottom-up methods provide better control of phase composition and morphology. Figure 2 shows electron microscopy images illustrating the morphology of MoS2 obtained by different methods where the difference between top-down methods and bottom-up methods is visible. While the bottom-up methods present defined nanosheets (Figure 2d–f), top-down methods create structures with irregular morphology (Figure 2a–c).

3.1. Top-Down Methods

One of the main top-down methods is mechanical exfoliation through the ‘Scotch-tape’ method. The basis of this method is to strip the MoS2 flakes using tapes. The ‘Scotch-tape’ method is considered a non-destructive method in which chemical reactions are not involved [10]. Although this method has a low yield, is time-consuming, and leads to a low quantity of product, it produces SLMoS2 in the form of flakes and is a simple process requiring accessible equipment. This technique also enables MoS2 with high quality and crystallinity in 2D [10,84]. Ball-milling is another mechanical exfoliation method. This method can be employed in both dry and wet states. Dry ball-milling can be used to greatly enhance the electrochemical and catalytic properties of bulk MoS2 [85]. However, wet ball-milling leads to the increased production of FLMoS2 when compared to dry ball-milling [10]. According to Tayyebi et al. [86], the pre-functionalization of the materials can weaken the force interacting between the layers, facilitating exfoliation during the ball-milling process.
Another top-down method is liquid-phase exfoliation. In this approach, the bulk material is dispersed in a solvent and then exfoliated via mechanical processes, such as ultrasonication [87]. This method enables the high-yield creation of 2D flakes in suspension as a scalable production [88,89]. However, after long periods of inactivity or after taking the solvent out, the nanosheets could restack back into bulk form. The addition of a polymer may stabilize the MoS2 nanosheets, thereby maintaining the properties of exfoliated nanosheets. For instance, Deepak et al. [77] added the polymer poly(vinylidene fluoride-co-hexafluoropropylene) (PVDF-HFP) and used dimethylformamide (DMF) as a solvent in the liquid-phase exfoliation method. Due to its surface energy, water is not considered a suitable solvent for this method. However, hot water has already been employed to exfoliate MoS2, with satisfactory results [87]. The liquid-phase exfoliation can be assisted by ion intercalation to enhance the yield. Ions such as lithium can penetrate the layers of MoS2, weakening the interlayer forces and facilitating the detachment of the layers, allowing the production of SLMoS2. However, this method decreases the semiconducting properties of the resulting MoS2 [37].

3.2. Bottom-Up Methods

The chemical vapor deposition (CVD) method enables a controllable route to decompose Mo and S precursors and assemble them into nanosheets [10]. The MoS2 nanosheets obtained by CVD are thin, of high quality and high purity, and present limited defects as 2D materials [90]. This method is also indicated for producing SLMoS2 in large quantities at a low cost. However, it presents low control in terms of target products [91] and lacks good reproducibility since it depends on various parameters, such as temperature, pressure, and reaction time, among others [81]. The possibility of forming both SLMoS2 and FLMoS2 zones is also a limitation.
An alternative bottom-up approach that is frequently used is the hydrothermal method. In this method, the MoS2 crystallizes from aqueous or organic solutions processed at high pressure and temperature [10]. This method achieves high yield, controllable size, and uniform thickness [10]. Although the hydrothermal method is widely used, the mechanism is unclear as the reaction occurs in a closed environment, an autoclave [10]. Different precursors can be used in this method. Khabiri et al. [92] used the hydrothermal method to prepare composites containing MoS2 nanoflower structures, consisting of adding ammonium molybdate, thiourea, and water to an autoclave and setting the temperature to 180 °C for 20 h. Other authors used a similar procedure, but added ammonium tetrathiomolybdate, ethanol, and water to an autoclave heated at 200 °C for 10 min [93].
In the hydrothermal method, the morphology and properties of MoS2 can be modified by adjusting the experimental synthesis conditions. Luo et al. [83] observed different morphologies, depending on the S:Mo molar ratio of the starting materials of molybdenum oxide (MoO3) and potassium thiocyanate (KSCN). Low ratios led to the formation of nano-flower structures (Figure 3a), while at higher ratios (Figure 3b), only nanosheet structures were observed. The temperature and time of the synthesis can also influence the morphology. It was found that at a lower temperature (150 °C) for 25 h, the resulting morphology was like a coral (Figure 3c); when increasing either the time or temperature, the dominant morphology was a flower-like structure (Figure 3d), with size increasing over the synthesis time. At higher temperatures and longer reaction times (240 °C for 47 h), the observed structure consisted of large nanosheets, showing the largest surface area. Other structures, such as core-shell or hollow structures, have been produced, aiming to enhance the properties of MoS2 [94,95]. Wu et al. [95] produced 2-µm diameter hollow MoS2 microspheres via the hydrothermal method and used this material to adsorb the organic contaminant, methyl orange.

3.3. MoS2 Modification Methods

The properties of MoS2 can be modified by expanding the interlaying spacing, a strategy that has been applied in the fields of batteries and photocatalysis [96]. Ai et al. [97] produced a widened defect-rich nano MoS2 (W-DR-N-MoS2) via the hydrothermal method to adsorb mercury (Hg2+). The interlayer spacing could be controlled by the synthesis temperatures. The produced MoS2 consisted of layers of MoS2 with a wider space between them (d = 9.4 Å against d = 6.15 Å for MoS2 powder); empty spaces in the layers also characterized this material. Both aspects improved the adsorption capability by enabling adsorption between the layers.
The combination of MoS2 with other components can lead to materials with improved adsorption capacity. For example, kaolin was used as a substrate for MoS2 growth via the hydrothermal method. The resulting composite showed increased pore volume, a larger number of adsorption sites, and a consequently greater Pb(II) adsorption efficiency than their counterparts [98]. On the other hand, using montmorillonite (MMT) decreased the hydrophobicity of MoS2, increasing the dispersibility in water and improving the adsorptive removal of Hg(II) [99]. The growth of Fe3O4 nanoparticles onto MoS2 nanosheets resulted in MoS2 composites with enhanced conductivity and magnetic behavior [100,101].
Doping MoS2 with heteroatoms can also modify the MoS2 properties by regulating the electronic structure and conductivity, regulating the interlaying spacing, and increasing the number of active sites [102,103,104]. This type of modification has been shown to be effective in photocatalytic hydrogen production and battery fields [102,103,104]. Doping with metallic elements may lead to undesirable consequences, such as the reduction in stability of MoS2 by promoting the formation of the MoS3 phase. In contrast, using non-metallic elements promotes an increased number of active sites and conductivity [103]. Xin et al. [102] synthesized P-doped MoS2 for photocatalytic hydrogen production. P-doped MoS2 exhibited broader spectra absorption and a hydrogen production rate 2.8 times higher than the MoS2 counterpart.
Phase engineering is an approach in which several phases are combined to obtain optimal MoS2 properties. As mentioned above, a combined 1T/2H phase achieved better results in the photodegradation of methyl orange due to the increased conductivity of the 1T phase and the defects present in this phase [69]. An identical MoS2 phase combination was used in the photodegradation of potassium hexyl dithiocarbonate, which is an organic pollutant in mine wastewater [105].
An alternative strategy to modify the MoS2 properties is defect engineering, which consists of creating defect sites in MoS2 layers to act as active sites for catalysis [69,97]. Luo et al. [106] produced MoS2 with S defects using a ball-milling method in ascorbic acid, an organic acid-reducing agent. The resulting S defect-rich ultrathin MoS2 nanosheets were used to photo-reduce Cr(VI). This procedure leads to the formation of two types of S-defects: point defects that may act as recombination centers for photogenerated carriers, and stripping defects that promote the separation of photogenerated electron-hole pairs. According to the authors, combining both types of defects led to an enhancement of photocatalytic performance.
In the normal MoS2, only the edges sites have active regions with catalytic activity. One way to enhance this property is via the construction of vertically oriented MoS2 nanosheets (V-MoS2), which exhibit higher amounts of edge sites, enhanced longitudinal in-plane carrier transport, and stronger light adsorption [107,108]. Liu et al. [107] produced V-MoS2 by means of CVD induced by a TiO2 buffer layer, for application in photodetectors. Cui et al. [108] successfully applied a vertical growth MoS2 construction method in graphene sheets to increase the electrochemical performance of lithium-sulfur batteries.

4. Adsorption Applications

This section reviews the application of MoS2-based nanomaterials to clean up emerging pollutants from water using adsorptive technologies. Table 1 provides an overview of several MoS2-based nanomaterials investigated for the adsorption of inorganic and organic contaminants.

4.1. Heavy Metal Species

Pollution with heavy metal species is a huge concern because it represents a high risk for the environment and for humans [8]. Heavy metal ions are toxic, non-biodegradable, persistent in the environment, and tend to accumulate in organisms, causing health issues [8]. Lead (Pb(II)) is one of the most common ions found in industrial wastewater and shows toxicity, even at low concentrations (≥15 µg/L in drinking water) [8]. Many problems, such as anemia, hypertension, reduced intelligence quotient, and immunotoxicity are related to Pb(II) exposure [8]. Mercury is another heavy metal that presents a high risk to human health and may cause kidney failure, severe pulmonary irritation, neurological disorders, and death [73]. Mercury exists in the form of several inorganic (e.g., metallic mercury or mercuric salts (Hg(II)) and organometallic species (e.g., methylmercury). The Hg(II) species can also affect the food chain once they can bioaccumulate [73]. Chromium, particularly chromium (VI), is hazardous to human health as well [112]. In this oxidation state, Cr(VI) can cause cancer and fetal malformations with only a concentration higher than 0.10 mg/L in drinking water [112].
As stated above, MoS2 has good adsorption capability for heavy metal ions due to the sulfur atoms on the surface. For this reason, MoS2 has been investigated in terms of the adsorptive removal of these pollutants from water. Table 1 lists the most relevant studies. Most of the studies focused firstly on Hg(II) removal, due to its high toxicity, and, secondly, on Pb(II) and Cr(VI) removal. For instance, Pirarath et al. [73] produced nanosheets of MoS2 via surfactant-assisted hydrothermal synthesis, using sodium dodecyl sulfate (SDS), and investigated the adsorption of Hg(II) in batch conditions. The negative charge of the S layers increased with SDS, which induced an electrostatic interaction between Hg(II) and MoS2. A higher adsorption capacity was achieved due to the excess negative charge of the S layer, compared to the nanosheets of MoS2 produced by other methods. Within 300 min, 0.5 g/L of the composite enabled 93% removal of Hg(II) in distilled water samples. Ai et al. [97] produced a widened defect-rich nano MoS2 (W-DR-N- MoS2), with many sulfur binding sites exposed due to enlarged interlayer spacing. With this modification, Hg(II) removal of 99.8% was achieved in just 5 min in distilled water samples, using 0.1 g/L of the material. The Hg(II) removal was also efficient in both natural water and industrial wastewater samples. The recovery of the adsorbent is economically important. Coupling a magnetic component to MoS2 has been investigated as a strategy to facilitate its separation; employing a magnetic field is an accessible method. Zhi et al. [109] produced an aerogel of MoS2 and GO with embedded Fe3O4 and Au NPs (Au/Fe3O4/MoS2CAs) to remove Hg(II). A removal performance of 100% was achieved in only 30 min with 0.5 g/L of Au/Fe3O4/MoS2CAs, which is a great achievement for Hg(II) adsorption. The reuse of this composite for up to 10 cycles only slightly decreased the efficiency of the adsorption. Wang et al. [111] used flower-like MoS2 decorated with Fe3O4 to investigate the simultaneous removal of Hg(II) and Pb(II) from distilled water, along with Pb(II) removal in real battery wastewater and soil samples, achieving satisfactory results. Yuan et al. [98] produced a MoS2-kaolin composite for Pb(II) adsorption, reaching a 99.9% removal rate in 40 min with 1.6 g/L of composite. According to the authors, using kaolin induces uniform MoS2 growth and a higher pore volume, leading to more adsorption sites. Regarding the Cr(VI) removal, Chen et al. [112] produced MoS2/lignin-derived carbon (MoS2/LDC) and achieved 99.4% removal of Cr(VI) in distilled water samples after 30 min, using just 0.10 g/L of composite.
It is essential to evaluate the selectivity of the adsorbents toward the target pollutant because in natural water or wastewater, there are other species that can compete for sorption sites. In this context, Zhi et al. [109] evaluated the selectivity of the magnetic Au/Fe3O4/MoS2CAs composites toward Hg(II) capture via a mixture of a large number of ions in a synthetic water sample. Despite the presence of several ions (Na(I), K(I), Ag(I), Ca(II), Ba(II), Mg(II), Fe(II), Cu(II), Zn(II), Ni(II), Cd(II), Al(III), Pb(II), Cr(III) and Hg(II)) the most adsorbed ion was Hg(II). All the other ions did not show relevant adsorption. The reason for this selectivity may be related to the HSAB principle, i.e., soft acids such as Hg(II) establish strong interactions with soft bases such as sulfur [111,115].
After adsorption tests, the composite with the adsorbed heavy metal ion can be separated from the solution and treated, aiming for a reduction in the heavy metal ions, which precipitate and desorb from the catalyst [111,112].
Besides the heavy metal species, MoS2-based materials have been employed for the capture of noble metal species, namely, silver and gold ions. The recovery of these metals from industrial effluents and wastewater is important, not only to prevent pollution but also because of their economic value [28,74]. Zeng et al. [74] produced a 2D MoS2 to remove Ag(I) from distilled water samples. The mechanism of adsorption was investigated using X-ray photoelectron spectroscopy (XPS) and DFT calculations, confirming a strong interaction between MoS2 and Ag(I) by the complexation of Ag-S and Ag-O through hybridization between the d orbital of Ag and the p orbital of S or O [74]. The recovery of gold ions (Au(III)) has also been investigated [28]. Using chitosan-coated MoS2 (0.28 g/L dose), Zhao et al. [28] achieved 98.9% ion removal in 300 min. The ion-selectivity of chitosan-coated MoS2 in wastewater containing Au(III), Cu(II), Mg(II), Ni(I), Li(I), and Zn(II) species was evaluated. It was found that the biosorbent presented the highest selectivity to the Au(III) ions and the lowest affinity to Zn(II) ions. After adsorption, desorption enabled the recovery of noble metal ions for further use.

4.2. Organic Contaminants

The adsorption of organic contaminants using MoS2-based materials was less widely investigated. MoS2-based membranes for organic contaminant removal are beyond the scope of this work and have recently been reviewed elsewhere [13]. Most research papers that present results of the adsorption of organic contaminants have, as their primary focus, the study of the photodegradation of these contaminants. Table 1 lists several studies of the adsorption of organic contaminants. MoS2 nanosheets produced by the hydrothermal method removed nearly 82.3% of Bisphenol A (BA), which is an endocrine disruptor, after 120 min with a sorbent dosage of 1.5 g/L [83]. Reduced graphene oxide-MoS2 (rGO-MoS2) composites successfully removed the ofloxacin antibiotic [7]. Despite the low surface area (17 m2/g), 95% removal was achieved after 240 min with 0.35 g/L of adsorbent. The adsorption capacity in real wastewater samples was also investigated. However, the removal drastically decreased after composite reuse, and only 4% was removed after 10 cycles. Gao et al. [113] produced a composite with a metal–organic framework (UiO-66/MoS2) and report a comprehensive study on the adsorption and photocatalytic degradation of lomefloxacin. The beta-blockers, atenolol and acebutolol, were also adsorbed successfully in distilled water with MoS2/montmorillonite (MoS2/MTT) composites [114]. Regarding the adsorption of dyes, methyl orange was adsorbed by hollow MoS2 microspheres (h-MoS2) [95]. The adsorption equilibrium was achieved in just 10 s for MO, with a maximum adsorption capacity of 42 mg/g.
MoS2-based composites have also been used for the adsorption of various organic dyes. Song et al. [100] produced MoS2 decorated with Fe3O4 nanoparticles and tested this composite for the adsorption of Congo red (CR), methylene blue (MB), methylene green (MG), rhodamine B (RhB) and eosin Y (EY). Among these dyes, the composite showed higher adsorption of CR with an adsorption capacity of 71 mg/g, achieving the removal of about 65.2% in 2 min, which increased to 71% after 120 min with 1 g/L of the composite.

4.3. Kinetic and Isotherm Studies

Table 1 presents a summary of relevant information on the kinetics and equilibrium adsorption tests employing MoS2-based materials. Overall, the kinetic model that better described the results was the pseudo-second-order (PSO) kinetic model, with an R2 value of between 0.901–1.000, indicating a good fit of this model to the adsorption data of both the metal species and organic contaminants. Regarding the equilibrium description, the Langmuir model better described the adsorption data (R2 = 0.86–1.0). However, the equilibrium adsorption of the antibiotic lomefloxacin and MO dye was better described by the Freundlich isotherm model, indicating non-monolayer formation [83,115]. Wang et al. [111] used flower-like MoS2 decorated with Fe3O4 nanoparticles in the removal of Hg(II) and Pb(II) from an aqueous environment. The maximum adsorption capacity estimated by the Langmuir model was 428.9 mg/g for Hg(II) and 263.6 mg/g for Pb(II), under optimized conditions (pH 5, 25 °C, 0.8 g L−1 sorbent dosage). The adsorption kinetics were well described by the PSO model and the good fitting of the Langmuir isotherm model was in line with monolayer formation. The favorable adsorption capacity, selectivity, and recyclability originated from the strong interaction between S and the heavy metal ions. The MoS2-clinoptilolite composite removed nearly 100% (99.8%) of Pb(II) from aqueous solution (pH 6, 25 °C, starting concentration of 50 mg L−1, 1.5 g L−1 sorbent dosage) after 90 min [8]. This adsorption was ascribed to the interaction of Pb(II), not only with the S layers of the MoS2 but also with the -OH and -COOH functional groups of clinoptilodite. The thermodynamic parameters were calculated and indicated that the adsorption was spontaneous and exothermic in nature.
Looking at the equilibrium adsorption values (qe) of the metal species in Table 1, the highest value found was 3435 mg/g (R2 = 0.991) for the adsorption of Au(III) at an optimal pH of 5, using chitosan-coated MoS2 (CS/MoS2) crosslinked with glutaraldehyde. The value was estimated by the Langmuir model and was close to the experimental value (3109 mg/g) [28]. The adsorption capacity increased and the adsorption isotherm changed from multilayer to monolayer when the content of MoS2 increased. The highest adsorption capacity was observed for a mass ratio of 1:2 (CS:MoS2), at 35 °C, after 300 min. The high efficiency of these materials to adsorb Au(III) was ascribed to electrostatic interaction, in the form of tetrachloroaurate ions (AuCl4-) and then to complexation by the sulfur- and nitrogen-containing functional group. The materials also showed outstanding selectivity for gold ions in the presence of coexisting ions and demonstrated attractive reusability after four cycles.
Concerning the organic contaminants, the highest adsorption capacity was observed for atenolol, using MoS2/MTT as an adsorbent (qe = 146 mg/g) and close to the value estimated by the Langmuir model (132 mg/g) [114]. The adsorption kinetics could be described using the PSO model. The interactions leading to the adsorption included van der Waals interactions and hydrogen bonding between the hydroxyl groups of the atenolol and the sorbent. The quantum chemical calculations results were in line with the lower adsorption capacity observed for acetobutolol using the same adsorbent.
Other parameters and isotherm models have been used to characterize the adsorption, such as the distribution coefficient, which expresses the affinity of a compound to the adsorbent, and Temkin isotherm, which assumes that the heat of adsorption does not remain constant during the adsorption process [74,110].

5. Photocatalytic Applications

Adsorption is an excellent approach to removing pollutants from wastewater. However, this technique does not provide a complete solution to the problem since the contaminants are not mineralized. In this sense, advanced oxidation processes (AOPs), such as photocatalytic oxidation, have been developed to promote the degradation of pollutants [116].
Due to the relatively small band gap of MoS2, visible-light irradiation can induce a catalytic chain reaction. For bulk MoS2, the band gap is ~1.2 eV, which means that MoS2 is activated with almost all the solar spectrum, while SLMoS2 shows a band gap of ~1.8 eV, this phase being activated by a radiation wavelength of <660 nm. However, as the bulk MoS2 has unsaturated Mo and S atoms at the edge, leading to an indirect and small band gap, this is insufficient to achieve photocatalytic reactions. For this reason, bulk MoS2 is not suitable for photocatalysis [117].
MoS2 has been combined with several components to improve its performance as a photocatalyst to degrade contaminants. Table 2 lists several recent studies on the photocatalytic degradation of pollutants in water. MoS2-based materials have been primarily employed in the photodegradation of dyes. Dye-related industries are reported as the major source of contaminated wastewater [118]. Dye-containing wastewater presents both chromaticity and toxicity [25,119]. Pharmaceuticals, antibiotics, and pesticides, among other chemical compounds, also represent emerging contaminants where photodegradation has been investigated, due to their toxicity and increasing detection in wastewater [119]. The same photocatalyst could photodegrade various contaminants at different rates. For example, Fu et al. [120] reported degradation rates above 70% for methylene blue (MB), methyl orange (MO), and Congo red (CR), using ZnO/MoS2 nanoarrays. In contrast, the degradation of rhodamine B (RhB) dye was only around 8%. Regarding the degradation of MB, Khabiri et al. [92] achieved 83% degradation after 1 min of irradiation with VIS light, that increased to 91% after 120 min using α-Fe2O3/MoS2 QDs, a composite comprising hematite (α-Fe2O3) and nanosized MoS2 (MoS2 quantum dots (QDs)). Figure 4 illustrates the proposed photocatalytic mechanism. The visible light irradiation excites the electrons (e) in the valence bands (VB) of α-Fe2O3 and MoS2QDs to the respective conduction bands (CB), leaving a hole (h+) in the VB. The CB and VB positions of MoS2QDs are above α-Fe2O3. Due to the potential energy difference, the photogenerated electrons at the CB of MoS2QDs transfer to the CB of α-Fe2O3, then migrate to the surface of the catalyst, which enables the creation of the reactive species as a superoxide anion radical (·O2) by the reaction with O2, and hydroxyl radicals (·OH) from reaction with water. The reactive species react with MB, thereby degrading it [92]. Most of the values for the degradation rate presented in Table 2 are above 70% for different contaminants, MoS2-based catalysts, and methodologies. For MB, the highest degradation rate was 97%, after 40 min of reaction using 1 g/L of MZO (MoS2/ZnO) [121]. A similar degradation rate was obtained in just 5 min using the ternary catalyst MoS2/BiFeO3/Ag3PO4 under ultraviolet-visible (UV-Vis) irradiation [119]. The same catalyst achieved a good degradation of MO and RhB, with 98% and 97% degradation, respectively, after 5 min of irradiation (Table 2) [119]. Regarding other contaminants from the antibiotic spectrum, tetracycline (20 mg/L) was successfully degraded up to 99% within 50 min under visible light, catalyzed with 0.4 g/L MoS2/Eu/B-g-C3N4 (MoS2/BEuCN) as the catalyst [122]. Ahamad et al. [123] studied the photocatalytic degradation of bisphenol-A (BA) using the composite g-C3N4/MoS2−PANI, achieving 93% of removal after 60 min under visible light. DFT calculations were performed to elucidate the photocatalytic degradation mechanism of BA. It was proposed that the aromatic ring would be one of the most likely sites for an attack by photo-generated radicals, which leads to several intermediate species and to the production of the final products, CO2 and H2O [123].
Although most photodegradation studies rely on organic contaminants, MoS2-based photocatalysts were also employed to reduce Cr(VI) to Cr(III) [120,134]. As mentioned before, Cr(VI) is a highly toxic form of chromium. By reducing it to Cr(III), the toxicity decreases, and the chromium can be precipitated and removed [144]. Via photo-reduction, 85% Cr(VI) removal was achieved using MoS2/BiOBr/CF under visible light after 2 h, and 75% Cr removal rate using Ni foam with ZnO/MoS2 after 25 min under UV-VIS light [120,134]. Shi et al. [134] tested the effect of pH in this photocatalytic reaction and observed the highest removal (91%) at pH = 3 and the lowest removal (29%) at pH = 11. This behavior was ascribed to different chromium speciation according to pH. At pH < 7, the predominant species is the dichromate anion (Cr2O72), which is more likely to be reduced to Cr(III) than the chromate anion (CrO42−), which predominates at pH > 7. In addition, using MoS2/TiO2/graphene, Chen et al. [137] achieved an 88% removal of Pb(II) in an aqueous solution in just 60 min under UV-Vis light. The removal was higher in an acidic environment (pH < 7) because the dominant species at those pH values (Pb2+(H2O)6) is more reactive than the other Pb(II) species, due to the more positive charge and smaller hydration radius.
In photocatalysis, the goal is to achieve the total mineralization of the pollutants. This means fully converting the contaminants to a gaseous form, water, salts, and minerals [145]. However, several photodegradation reactions may lead to toxic intermediate compounds, which can harm the environment [146]. The total organic carbon (TOC) quantification method can be employed to evaluate the mineralization of a contaminant achieved by photocatalysis. Table 3 shows a relatively high mineralization rate, with more than 54% of the mineralization provided by MoS2-based photocatalysts. This means that most of the contaminant molecules present in the solution (>50%) were fully degraded, while part may still be the original contaminant (when the removal rate is less than 100%) or is converted into some intermediate compound. For example, 100% degradation of Rhodamine B was achieved with the BiOI/MoS2 (BMS-8) catalyst. However, only 78% of the initial amount of the dye was fully mineralized, leaving 23% in the form of intermediate compounds [3]. Overall, the results show the promising potential of MoS2-based catalysts to photodegrade and mineralize the organic contaminants in wastewater systems.

6. MoS2-Based Nanocomposites

Although nano-MoS2 shows good efficiency as a photocatalyst for removing several contaminants, it can be enhanced when in combination with other components in the form of nanocomposites. Table 4 lists several works comparing the photodegradation provided by bare nano-MoS2 and MoS2-based nanocomposites. Overall, MoS2-based nanocomposites present a higher specific area when compared to bare nano-MoS2. The increase in the specific surface area of the catalyst can contribute to an enhanced removal rate once a higher exposed specific area favors the contaminant-catalyst contact [142].
Overall MoS2-based composites presented better adsorptive performance than bare MoS2. The creation of defects and increasing interlayer spacing were used to enhance the adsorption capacity of Hg(II) from 36 mg/g to 2563 mg/g [97]. The introduction of montmorillonite (MMT) enhanced the water dispersibility of MoS2 by decreasing the hydrophobicity, which improved the adsorption capacity [99]. For instance, MoS2/MMT composites adsorbed nearly 1120 mg/g of Hg(II), which is markedly higher than the value observed for bare MoS2 (937 mg/g). MoS2/MMT composite was also used to adsorb the beta-blockers atenolol and acebutolol, achieving good results [114]. Kaolin has been used as a substrate, providing a uniform growth of MoS2 and well-distributed active sites, which improved the removal of Pb(II) from 55% in bare MoS2 to 89% in MoS2-kaolin [98].
The combination of MoS2 with semiconducting phases has been explored, to improve photocatalytic activity. Quan et al. [121] coupled MoS2 nanoflowers to ZnO nanoparticles and observed a considerable increase in the degradation rate of methylene blue compared with bare MoS2, from 44% to 97%, when submitted to visible light (λ > 420 nm). Introducing the semiconductor ZnO resulted in an increased photoinduced electron transfer rate, leading to higher photocatalytic activity. ZnO also presents attractive characteristics, such as low cost, low toxicity, high chemical stability, and strong photosensitivity [121]. Mahalakshmi et al. [127] produced core@shell TiO2@MoS2 heterojunction composites via the one-step hydrothermal method. The photocatalytic activity of the resulting composite toward 4-nitrophenol was enhanced compared to the MoS2 and TiO2 used separately, due to a reduced band gap and efficient separation of the photogenerated electron-hole pairs. A similar strategy was reported for coupling MoS2 and CuO, with improved separation of electron-hole pairs [142]. A Z-scheme heterojunction was produced that promoted the separation of the photogenerated carriers and may also increase the specific area and, consequently, the number of active sites [142]. BiOI nanoplates were also combined with MoS2 nanosheets [3]. ZnS was combined with MoS2 to increase the efficiency of photocarrier generation, increasing the removal of crystal violet dye using visible light (λ > 420 nm) [132]. Composites of MoS2 with UiO-66 resulted in an enhanced specific area and in more active sites being available [113]. Several studies report the incorporation of magnetic nanostructures (e.g., magnetite—Fe3O4) to enhance the photocatalytic activity of MoS2 [25,135]. This approach facilitates the separation of the nanocatalyst from the treated solution and increases the efficiency of the transport of photogenerated electrons, leading to a higher removal rate, even in the case of a lower specific area compared to bare MoS2 [25].

7. Reuse of MoS2-Based Composites for Adsorption and Photocatalysis

After interacting with the contaminated solution, the catalysts and adsorbents should then be collected and reused to reduce the quantity of produced materials and the associated costs. Importantly, the composite should maintain good adsorption/photoactivity when reused. Reusability studies have been performed to verify the efficiency of the composite after several cycles. At the end of each removal experiment, the sorbent/photocatalyst is collected through centrifugation or filtration and is washed off with water and ethanol or water alone, and then dried [121,124]. In this context, magnetic nanomaterials are advantageous because they are easily recovered using magnetic separation, which requires less energy [25,100,111,135].
Table 5 compares the evolution of the degradation rate via adsorption and photodegradation after several testing cycles. The MoS2-based composites showed good adsorption when comparing the first and the last cycles [99,110,111,112]. The Au/MoS2 composites also presented good adsorption after four consecutive cycles in a solution containing multiple ions, demonstrating their great potential for heavy metal ion removal [28]. Zhi et al. [109] reported a removal rate of > 95% after 10 reusing cycles for the adsorption of Hg(II). Several other studies report minor differences in adsorption capacity at the first and last adsorption cycles, regardless of the contaminant. Conversely, a marked decrease in the adsorption of the antibiotic ofloxacin was observed when the rGO-MoS2 composite was reused [7]. The removal decreased from 92% to just 4% after 8 cycles, but the authors do not offer a possible explanation [7].
The reuse of MoS2-based photocatalysts has also been investigated. However, often, the papers do not give detailed results of the reuse studies. Overall, the photocatalyst maintains good photocatalytic activity after several cycles. Yet, in some cases, significant decreases in the photodegradation rate were observed due to the oxidation of the photocatalyst [126].

8. Application of MoS2-Based Composites in Real Environmental Samples

So far, most studies have only focused on treating singular contaminant solutions in synthetic aqueous samples. The number of reports testing MoS2 and composites in water decontamination in more realistic conditions is still scarce. MoS2/Fe3O4 composite was employed as an adsorbent to treat effluents from lead-acid battery factories, reaching a Pb(II) removal of 99.6% and showing acceptable Pb(II) removal in contaminated soil [111]. Regarding the adsorption of organic contaminants, Zeng et al. [147] used MoS2-decorated biochar to remove the antibiotic tetracycline hydrochloride from river and tap water. The adsorption capacity was found to be higher in environmental water samples when compared to synthetic solutions, prepared in deionized water and used under the same conditions. Furthermore, the composite could be reused for 5 cycles with only a minor decrease in the adsorption capacity. Conversely, the adsorption capacity of the antibiotic ofloxacin by the rGO-MoS2 composites in river water was lower than in deionized water [7]. The effect was ascribed to the presence of cations, natural and synthetic organic chemicals that compete with ofloxacin for adsorption, leading to a decrease in the adsorption capacity. Huang et al. [148] realized that the photocatalytic activity of MoS2 microspheres in degrading thiobencarb in lake or river water decreased by 20% compared to its performance in spiked deionized water, under identical reaction conditions. The presence of anions in non-deionized water acting as scavengers was identified as a cause for the decrease in photocatalytic activity. Nevertheless, the authors considered this material a potential catalyst for removing pollutants. Chandrabose et al. [149] assessed the photocatalytic efficiency of MoS2/TiO2 composite in a solution comprising a mixture of anionic dyes (rhodamine B and methyl orange) and cationic dyes (methylene blue and crystal violet). The composite removed 100% of cationic dyes and 70% of the anionic dyes in the first stage of adsorption in the dark, while the remaining 30% of anionic dyes were photodegraded entirely after 3.5 h in a second stage under UV-Vis light irradiation. The composite showed higher adsorptive affinity to the cationic dyes compared to anionic dyes. Increasing contaminant concentrations resulted in less adsorptive removal at the first stage, which was limited by the surface area of the composite. The results of the study show the potential of using this MoS2-based composite for the removal of dyes produced by textile industries.

9. Environmental Impact of MoS2-Based Composites

Water remediation studies employing MoS2-based composites have been focused on the evaluation and enhancement of the composites’ removal capacity. The assessment of the ecotoxicity of the materials has been less widely investigated, although it is a fundamental matter. Nanomaterials have emerged as a promising tool for water remediation, but, due to their high surface area and chemical activity, if accidentally released into the environment, they may constitute a risk [150]. Depending on their chemical properties and composition, nanomaterials can be modified by oxidation, sulfidation, aggregation, and deposition in the natural environment [151]. In terms of the properties of nano-MoS2, this material is considered chemically stable even in environmental conditions [75]. Due to its low solubility in water systems, it is also considered a persisting compound. Nano-MoS2 may persist in the environment and the actual risks are still unknown. The nanosheets of MoS2 showed antibacterial activity against Escherichia coli and Staphylococcus aureus and low cytotoxicity for Vero cells [19]. Cicuendez et al. [152] also observed good viability of mouse and human cells when exposed to MoS2 flakes. A more comprehensive study on the toxicity of bulk and MoS2 nanosheets was conducted to address the toxicity in various aquatic species of different taxonomic groups [34]. The EC50 was determined for four species: Vibrio fischeri, a marine Gram-negative photobacterium, Pseudokirchnerialla subcapitata, a freshwater microalga, Daphnia magna, a freshwater crustacean, and Spirodela polyrhiza, a freshwater duckweed. Bulk and nano MoS2 (0.05–2.00 g/L) were tested for a contact time ranging from 15 min to 72 h. It was found that bulk MoS2 was more toxic than MoS2 nanosheets for all the species tested; in some cases, nano MoS2 showed no toxicity for the tested conditions.

10. Conclusions and Perspectives

In summary, this review highlights the advantages of MoS2 and MoS2–based nanomaterials for water treatment applications through adsorption and photocatalysis. Owing to the high surface area and visible light-responsive photocatalytic activity, we can expect growing interest in these materials for environmental applications in the near future. Several advances have been made in synthetic methods for the production of MoS2 nanomaterials with different structural features, namely, the crystal phase, morphology and layer spacing, which each create distinct physicochemical properties. The applications of MoS2 and composites in the adsorption and photodegradation of metal ion species and organic contaminants are thoroughly summarized. The modification of MoS2 and its combination with other components enhance those properties that increase removal efficiency and may also facilitate the separation of the catalyst from the solution. Despite the progress achieved, more studies are needed to rationally fabricate highly efficient MoS2-based photocatalysts, which represents an opportunity for further research efforts. The ecotoxicity of MoS2 is still not clear and requires further investigation. Although MoS2 presents high chemical stability, even in environmental aqueous systems, it also presents some toxicity to different organisms. More studies should be conducted to assess the risks of employing MoS2-based materials in water treatment.

Author Contributions

Conceptualization, L.O.A. and A.L.D.-d.-S.; methodology, L.O.A. and A.L.D.-d.-S.; investigation, L.O.A.; writing—original draft preparation, L.O.A.; writing—review and editing, L.O.A. and A.L.D.-d.-S.; supervision, A.L.D.-d.-S.; funding acquisition, A.L.D.-d.-S. All authors have read and agreed to the published version of the manuscript.

Funding

L. O. Amaral thanks the Fundação para a Ciência e Tecnologia (FCT) for Ph.D. grant number UI/BD/151138/2021. This work was developed within the scope of the project CICECO-Aveiro Institute of Materials, UIDB/50011/2020, UIDP/50011/2020, and LA/P/0006/2020, financed by national funds through the FCT/MCTES (PIDDAC).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yang, D.; Yang, Y.; Xia, J. Hydrological cycle and water resources in a changing world: A review. Geogr. Sustain. 2021, 2, 115–122. [Google Scholar] [CrossRef]
  2. Chen, X.; Shuai, C.; Wu, Y.; Zhang, Y. Understanding the sustainable consumption of energy resources in global industrial sector: Evidences from 114 countries. Environ. Impact Assess. Rev. 2021, 90, 106609. [Google Scholar] [CrossRef]
  3. Hao, L.; Ju, P.; Zhang, Y.; Zhai, X.; Sun, C.; Duan, J.; Su, Y.; Lu, Z.; Liao, D. Fabrication of hierarchical flower-like BiOI/MoS2 heterostructures with highly enhanced visible-light photocatalytic activities. Colloids Surf. A Physicochem. Eng. Asp. 2021, 610, 125714. [Google Scholar] [CrossRef]
  4. Liu, Y.; Ma, C.; Zhang, X.; Ngo, H.N.; Guo, W.; Zhang, M.; Zhang, D. Role of structural characteristics of MoS2 nanosheets on Pb2+ removal in aqueous solution. Environ. Technol. Innov. 2021, 22, 101385. [Google Scholar] [CrossRef]
  5. Shahid, M.K.; Kashif, A.; Fuwad, A.; Choi, Y. Current advances in treatment technologies for removal of emerging contaminants from water—A critical review. Coord. Chem. Rev. 2021, 442, 101385. [Google Scholar] [CrossRef]
  6. Soares, S.F.; Fernandes, T.; Trindade, T.; Daniel-da-Silva, A.L. Recent advances on magnetic biosorbents and their applications for water treatment. Environ. Chem. Lett. 2020, 18, 151–164. [Google Scholar] [CrossRef]
  7. Jaswal, A.; Kaur, M.; Singh, S.; Kansal, S.K.; Umar, A.; Garoufalis, C.S.; Baskoutas, S. Adsorptive removal of antibiotic ofloxacin in aqueous phase using rGO-MoS2 heterostructure. J. Hazard. Mater. 2021, 417, 125982. [Google Scholar] [CrossRef] [PubMed]
  8. Pandey, S.; Fosso-Kankeu, E.; Spiro, M.J.; Waanders, F.; Kumar, N.; Ray, S.S.; Kim, J.; Kang, M. Equilibrium, kinetic, and thermodynamic studies of lead ion adsorption from mine wastewater onto MoS2- clinoptilolite composite. Mater. Today Chem. 2020, 18, 100376. [Google Scholar] [CrossRef]
  9. Estrada, A.C.; Daniel-da-Silva, A.L.; Leal, C.; Monteiro, C.; Lopes, C.B.; Nogueira, H.I.S.; Lopes, I.; Martins, M.J.; Martins, N.C.T.; Gonçalves, N.P.F.; et al. Colloidal nanomaterials for water quality improvement and monitoring. Front. Chem. 2022, 10, 1011186. [Google Scholar] [CrossRef]
  10. Tan, C.; Cao, X.; Wu, X.J.; He, Q.; Yiang, J.; Zhang, X.; Chen, J.; Zhao, W.; Han, S.; Nam, G.H.; et al. Recent advances in ultrathin two-dimensional nanomaterials. Chem. Rev. 2017, 117, 6225–6331. [Google Scholar] [CrossRef]
  11. Darvishi, M.; Foroutan, M. Molecular investigation of water adsorption on MoS2 and graphene surfaces. J. Mol. Liq. 2017, 225, 1–10. [Google Scholar] [CrossRef]
  12. Singh, A.; Shirodkar, S.N.; Waghmare, U.V. 1H and 1T polymorphs, structural transitions and anomalous properties of (Mo,W) (S,Se)2 monolayers: First-principles analysis. 2D Mater. 2015, 2, 035013. [Google Scholar] [CrossRef]
  13. Liu, Y.; Zhao, Y.; Zhang, X.; Huang, X.; Liao, W.; Zhao, Y. MoS2-based membranes in water treatment and purification. Chem. Eng. J. 2021, 422, 130082. [Google Scholar] [CrossRef]
  14. Zhang, W.; Huang, J.K.; Chen, C.H.; Chang, Y.H.; Cheng, Y.J.; Li, L.J. High-gain phototransistors based on a CVD MoS2 monolayer. Adv. Mater. 2013, 25, 3456–3461. [Google Scholar] [CrossRef] [PubMed]
  15. Shan, J.; Li, J.; Chu, X.; Xu, M.; Jin, F.; Fang, X.; Wei, Z.; Wang, X. Enhanced photoresponse characteristics of transistors using CVD-grown MoS2/WS2 heterostructures. Appl. Surf. Sci. 2018, 443, 31–38. [Google Scholar] [CrossRef]
  16. Christinelli, W.A.; Shimizu, F.M.; Facure, M.H.M.; Cerri, R.; Oliveira Jr, O.N.; Correa, D.S.; Mattoso, L.H.C. Two-dimensional MoS2-based impedimetric electronic tongue for the discrimination of endocrine disrupting chemicals using machine learning. Sens. Actuators B Chem. 2021, 336, 129696. [Google Scholar] [CrossRef]
  17. Raadnui, S.; Mahathanabodee, S.; Tongsri, R.; Morakotjinda, M.; Wila, P. Comparison of dry sliding tribological behavior of SS 316L impregnated with MoS2 vs. h-BN solid lubricants: A statistical point of view. Wear 2021, 476, 203676. [Google Scholar] [CrossRef]
  18. Dawari, C.K.; Haq, I.; Mönkkönen, K.; Suvanto, M.; Saarinen, J.J. Reduced sliding friction on flat and microstructured metal injection molded (MIM) WC-Co hard metals with MoS2 composite lubricants. Tribol. Int. 2021, 160, 107020. [Google Scholar] [CrossRef]
  19. Kasinathan, K.; Marimuthu, K.; Murigesan, B.; Pandiyan, N.; Pandi, B.; Mahalingam, S.; Selvaraj, B. Cyclodextrin functionalized multi-layered MoS2 nanosheets and its biocidal activity against pathogenic bacteria and MCF-7 breast cancer cells: Synthesis, characterization and in-vitro biomedical evaluation. J. Mol. Liq. 2021, 323, 114631. [Google Scholar] [CrossRef]
  20. Sha, R.; Bhattacharyya, T.K. MoS2-based nanosensors in biomedical and environmental monitoring applications. Electrochim. Acta. 2020, 349, 136370. [Google Scholar] [CrossRef]
  21. Kumar, R.R.; Murugesan, T.; Dash, A.; Hsu, C.; Gupta, S.; Manikandan, A.; Anbalagan, A.k.; Lee, C.; Tai, N.; Chueh, Y.; et al. Ultrasensitive and light-activated NO2 gas sensor based on networked MoS2/ZnO nanohybrid with adsorption/desorption kinetics study. Appl. Surf. Sci. 2021, 536, 147933. [Google Scholar] [CrossRef]
  22. Zhang, Z.; Li, W.; Wang, R.; Li, H.; Yan, J.; Jin, Q.; Feng, P.; Wang, K.; Jiang, K. Crystal water assisting MoS2 nanoflowers for reversible zinc storage. J. Alloys. Compd. 2021, 872, 159599. [Google Scholar] [CrossRef]
  23. Zhang, L.; Mu, L.; Zhou, Q.; Hu, X. Solar-assisted fabrication of dimpled 2H-MoS2 membrane for highly efficient water desalination. Water Res. 2020, 170, 115367. [Google Scholar] [CrossRef]
  24. Peng, W.; Wang, W.; Qi, M.; Miao, Y.; Huang, Y.; Yu, F. Enhanced capacitive deionization of defect-containing MoS2/graphene composites through introducing appropriate MoS2 defect. Electrochim. Acta. 2021, 383, 138363. [Google Scholar] [CrossRef]
  25. Wang, Q.; Dong, S.; Zhang, D.; Yu, C.; Lu, J.; Wang, D.; Sun, J. Magnetically recyclable visible-light responsive MoS2@Fe3O4 photocatalysts targeting efficient wastewater treatment. J. Mater. Sci. 2018, 53, 1135–1147. [Google Scholar] [CrossRef]
  26. Zhu, Q.; Xu, Q.; Du, M.; Zeng, X.; Zhong, G.; Qiu, B.; Zhang, J. Recent progress of metal sulfide photocatalysts for solar energy conversion. Adv. Mater. 2022, 2202929. [Google Scholar] [CrossRef]
  27. Hasija, V.; Raizada, P.; Thakur, V.K.; Parwaz Khan, A.A.; Asiri, A.M.; Singh, P. An overview of strategies for enhancement in photocatalytic oxidative ability of MoS2 for water purification. J. Environ. Chem. Eng. 2020, 8, 104307. [Google Scholar] [CrossRef]
  28. Zhao, M.; Huang, Z.; Wang, S.; Zhang, L. Ultrahigh efficient and selective adsorption of Au(III) from water by novel Chitosan-coated MoS2 biosorbents: Performance and mechanisms. Chem. Eng. J. 2020, 401, 126006. [Google Scholar] [CrossRef]
  29. Yang, X.; Li, J.; Liang, T.; Ma, C.; Zhang, Y.; Chen, H.; Hanagata, N.; Su, H.; Xu, M. Antibacterial activity of two-dimensional MoS2 sheets. Nanoscale 2014, 6, 10126–10133. [Google Scholar] [CrossRef]
  30. Fan, J.; Li, Y.; Nguyen, H.N.; Yao, Y.; Rodrigues, D.F. Toxicity of exfoliated-MoS2 and annealed exfoliated-MoS2 towards planktonic cells, biofilms, and mammalian cells in the presence of electron donor. Environ. Sci. Nano. 2015, 2, 370–379. [Google Scholar] [CrossRef]
  31. Cheng, A.; Zhang, H.; Zhong, W.; Li, Z.; Tang, Y.; Li, Z. Enhanced electrochemical properties of single-layer MoS2 embedded in carbon nanofibers by electrospinning as anode materials for sodium-ion batteries. J. Electroanal. Chem. 2019, 843, 31–36. [Google Scholar] [CrossRef]
  32. Zhao, P.; Zheng, J.; Guo, P.; Jiang, Z.; Cao, L.; Wan, Y. Electronic and magnetic properties of Re-doped single-layer MoS2: A DFT study. Comput. Mater. Sci. 2017, 128, 287–293. [Google Scholar] [CrossRef]
  33. Gangwar, R.; Pandey, D.; Kancharlapalli, S.; Raychaudhuri, D.; Chakrabarti, A.; Banerjee, A.; Ghanti, T.K. Ab Initio Study of Adsorption of Fission Gas Atoms Xe and Kr on MoS2 Monolayer Functionalized with 3d Transition Metals. J. Phys. Chem. C. 2021, 125, 1493–1508. [Google Scholar] [CrossRef]
  34. Arefi-Oskoui, S.; Khataee, A.; Ucun, O.K.; Kobya, M.; Hanci, T.Ö.; Arslan-Alaton, I. Toxicity evaluation of bulk and nanosheet MoS2 catalysts using battery bioassays. Chemosphere 2021, 268, 128822. [Google Scholar] [CrossRef]
  35. Rani, A.; Singh, K.; Patel, A.S.; Chakraborti, A.; Kumar, S.; Ghosh, K.; Sharma, P. Visible light driven photocatalysis of organic dyes using SnO2 decorated MoS2 nanocomposites. Chem. Phys. Lett. 2020, 738, 136874. [Google Scholar] [CrossRef]
  36. Chang, M.J.; Cui, W.N.; Liu, J.; Wang, K.; Du, H.L.; Qiu, L.; Fan, S.M.; Luo, Z.M. Construction of novel TiO2/Bi4Ti3O12/MoS2 core/shell nanofibers for enhanced visible light photocatalysis. J. Mater. Sci. Technol. 2020, 36, 97–105. [Google Scholar] [CrossRef]
  37. Eda, G.; Yamaguchi, H.; Voiry, D.; Fujita, T.; Chen, M.; Chhowalla, M. Photoluminescence from chemically exfoliated MoS2. Nano Lett. 2011, 11, 5111–5116. [Google Scholar] [CrossRef]
  38. Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Single-layer MoS2 transistors. Nat. Nanotechnol. 2011, 6, 147–150. [Google Scholar] [CrossRef]
  39. Han, Y.; Chen, P.; Zhang, C.; Dong, J.; Liu, H. The buckling behavior of single-layer MoS2 sheets with kirigami-inspired structures under compression. Comput. Mater. Sci. 2021, 188, 110188. [Google Scholar] [CrossRef]
  40. Splendiani, A.; Sun, L.; Zhang, Y.; Li, T.; Kim, J.; Chim, C.Y.; Galli, G.; Wang, F. Emerging photoluminescence in monolayer MoS2. Nano Lett. 2010, 10, 1271–1275. [Google Scholar] [CrossRef]
  41. Han, B.; Hu, Y.H. MoS2 as a co-catalyst for photocatalytic hydrogen production from water. Energy Sci. Eng. 2016, 4, 285–304. [Google Scholar] [CrossRef] [Green Version]
  42. Szary, M.J.; Michalewicz, M.T.; Radny, M.W. Giant spin splitting induced by a symmetry-braking van der Waals interaction. Appl. Surf. Sci. 2019, 494, 619–626. [Google Scholar] [CrossRef]
  43. Kumar, S.; Mishra, T. Shock wave induced exfoliation of molybdenum disulfide (MoS2) in various solvents: All-atom molecular dynamics simulation. J. Mol. Liq. 2020, 314, 113671. [Google Scholar] [CrossRef]
  44. Zhao, X.; Ning, S.; Fu, W.; Pennycook, S.J.; Loh, K.P. Differentiating polymorphs in molybdenum disulfide via electron microscopy. Adv. Mater. 2018, 30, 1802397. [Google Scholar] [CrossRef]
  45. Szary, M.J. Al doped MoS2 for adsorption-based water collection. Appl. Surf. Sci. 2020, 529, 147083. [Google Scholar] [CrossRef]
  46. Ramsdell, L.S. Studies on silicon carbide. Am. Mineral. 1947, 32, 64–82. [Google Scholar]
  47. Krishnan, U.; Kaur, M.; Singh, K.; Kumar, M.; Kumar, A. A synoptic review of MoS2: Synthesis to applications. Superlattices Microstruct. 2019, 128, 274–297. [Google Scholar] [CrossRef]
  48. Wang, Z.; Mi, B. Environmental Applications of 2D Molybdenum Disulfide (MoS2) Nanosheets. Environ. Sci. Technol. 2017, 51, 8229–8244. [Google Scholar] [CrossRef]
  49. Singh, A.K.; Kumar, P.; Late, D.J.; Kumar, A.; Patel, S.; Singh, J. 2D layered transition metal dichalcogenides (MoS2): Synthesis, applications and theoretical aspects. Appl. Mater. Today 2018, 13, 242–270. [Google Scholar] [CrossRef]
  50. Evans, P.E.; Komesu, T.; Schwier, E.F.; Kumar, S.; Shimada, K.; Dowben, P.A. The band shifts in MoS2(0001) and WSe2(0001) induced by palladium adsorption. J. Phys. Condens. Matter. 2020, 32, 465001. [Google Scholar] [CrossRef]
  51. Yun, W.S.; Han, S.W.; Hong, S.C.; Kim, I.G.; Lee, J.D. Thickness and strain effects on electronic structures of transition metal dichalcogenides: 2H-MX2 semiconductors (M = Mo, W, X = S, Se, Te). Phys. Rev. B Condens. Matter. Mater. Phys. 2012, 85, 033305. [Google Scholar] [CrossRef]
  52. Tang, Q.; Jiang, D.E. Mechanism of hydrogen evolution reaction on 1T-MoS2 from first principles. ACS Catal. 2016, 6, 4953–4961. [Google Scholar] [CrossRef]
  53. Liu, Q.; Fang, Q.; Chu, W.; Wan, Y.; Li, X.; Xu, W.; Habib, M.; Tao, S.; Zhou, Y.; Liu, D.; et al. Electron-doped 1T-MoS2 via interface engineering for enhanced electrocatalytic hydrogen evolution. Chem. Mater. 2017, 29, 4738–4744. [Google Scholar] [CrossRef]
  54. Hu, T.; Li, R.; Dong, J. A new (2 × 1) dimerized structure of monolayer 1T-molybdenum disulfide, studied from first principles calculations. J. Chem. Phys. 2013, 139, 174702. [Google Scholar] [CrossRef]
  55. Wu, M.H.; Li, L.; Liu, N.; Wang, D.J.; Xue, Y.C.; Tang, L. Molybdenum disulfide (MoS2) as a co-catalyst for photocatalytic degradation of organic contaminants: A review. Process Saf. Environ. Prot. 2018, 118, 40–58. [Google Scholar]
  56. Kour, P.; Deeksha; Yadav, K. Electrochemical performance of mixed-phase 1T/2H MoS2 synthesized by conventional hydrothermal v/s microwave-assisted hydrothermal method for supercapacitor applications. J. Alloys. Compd. 2022, 922, 166194. [Google Scholar] [CrossRef]
  57. Mishra, S.; Maurya, P.K.; Mishra, A.K. 2H–MoS2 nanoflowers based high energy density solid state supercapacitor. Mater Chem Phys. 2020, 255, 123551. [Google Scholar] [CrossRef]
  58. Nie, C.; Yin, M.; Zhao, Y.; Zhao, C.; Zhang, B.; Song, X.; Yi, X.; Zhang, Y.; Luo, L.; Wang, S. Tailoring the fluorescent and electronic properties of 2H-MoS2 by step-by-step functionalization. J. Phys. Chem. C 2021, 125, 25739–25748. [Google Scholar] [CrossRef]
  59. Li, F.; Shen, T.; Xu, L.; Hu, C.; Qi, J. Strain improving the performance of a flexible monolayer MoS2 photodetector. Adv Electron Mater. 2019, 5, 1900803. [Google Scholar] [CrossRef]
  60. Bai, H.; Wu, Q.; Ai, H.; Liu, D.; Feng, J.; Ang, L.K.; Lu, Y.; Yang, M.; Pan, H. Interlayer-incorporation of MoS2 (TM-MoS2) to achieve unique magnetic and electronic properties for spintronics. Adv. Electron. Mater. 2022, 2200209. [Google Scholar] [CrossRef]
  61. Panasci, S.E.; Koos, A.; Schilirò, E.; Di Franco, S.; Greco, G.; Fiorenza, P.; Roccaforte, F.; Agnello, S.; Cannas, M.; Gelardi, F.M.; et al. Multiscale investigation of the structural, electrical and photoluminescence properties of MoS2 obtained by MoO3 sulfurization. Nanomaterials 2022, 12, 182. [Google Scholar] [CrossRef] [PubMed]
  62. Pradhan, G.; Sharma, A.K. Temperature controlled 1T/2H phase ratio modulation in mono- and a few layered MoS2 films. Appl. Surf. Sci. 2019, 479, 1236–1245. [Google Scholar] [CrossRef]
  63. Silambarasan, K.; Archana, J.; Harish, S.; Navaneethan, M.; Ganesh, R.S.; Ponnusamy, S.; Muthamizhchelvan, C.; Hara, K. One-step fabrication of ultrathin layered 1T@2H phase MoS2 with high catalytic activity based counter electrode for photovoltaic devices. J. Mater. Sci. Technol. 2020, 51, 94–101. [Google Scholar]
  64. Tan, D.; Willatzen, M.; Wang, Z.L. Prediction of strong piezoelectricity in 3R-MoS2 multilayer structures. Nano Energy 2018, 56, 512–515. [Google Scholar] [CrossRef]
  65. Strachan, J.; Masters, A.F.; Maschmeyer, T. 3R-MoS2 in review: History, status, and outlook. ACS Appl. Energy Mater. 2021, 4, 7405–7418. [Google Scholar] [CrossRef]
  66. Shi, J.; Yu, P.; Liu, F.; He, P.; Wang, R.; Qin, L.; Zhou, J.; Li, X.; Zhou, J.; Sui, X.; et al. 3R MoS2 with broken inversion symmetry: A promising ultrathin nonlinear optical device. Adv. Mater. 2017, 29, 1701486. [Google Scholar] [CrossRef] [PubMed]
  67. Anghel, S.; Chumakov, Y.; Kravtsov, V.; Volodina, G.; Mitioglu, A.; Plochocka, P.; Sushkevich, K.; Mishina, E.; Kulyuk, L. Site-selective luminescence spectroscopy of bound excitons and local band structure of chlorine intercalated 2H- and 3R-MoS2 polytypes. J. Lumin. 2016, 177, 331–336. [Google Scholar] [CrossRef]
  68. Chen, J.; Zhao, X.; Grinblat, G.; Chen, Z.; Tan, S.J.R.; Fu, W.; Ding, Z.; Abdelwahab, I.; Li, Y.; Geng, D.; et al. Homoepitaxial growth of large-scale highly organized transition metal dichalcogenide patterns. Adv. Mater. 2018, 30, 1704674. [Google Scholar] [CrossRef]
  69. Saber, M.R.; Khabiri, G.; Maarouf, A.A.; Ulbricht, M.; Khalil, A.S.G. A comparative study on the photocatalytic degradation of organic dyes using hybridized 1T/2H, 1T/3R and 2H MoS2 nano-sheets. RSC Adv. 2018, 8, 26364–26370. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Chen, J.; Cao, J.; Zhou, J.; Wang, W.; Zhang, Y.; Liu, X. Computational screening for enhanced hydrogen sensing by doped-2H and pristine-1T MoS2. Chem. Phys. Lett. 2020, 749, 137450. [Google Scholar] [CrossRef]
  71. Ma, J.; Zhang, Q.; Yang, J.; Feng, S.; Lei, M.; Quhe, R. Computational study of phase engineered transition metal dichalcogenides heterostructures. Comput. Mater. Sci. 2018, 142, 129–134. [Google Scholar] [CrossRef]
  72. Xiao, Y.; Tan, M.; Li, Z.; He, L.; Gao, B.; Chen, Y.; Zheng, Y.; Lin, B. Ethylenediamine-assisted phase engineering of 1T/2H–MoS2/graphene for efficient and stable electrocatalytic hydrogen evolution. Int. J. Hydrog. Energy 2021, 46, 11688–11700. [Google Scholar] [CrossRef]
  73. Pirarath, R.; Shivashanmugam, P.; Syed, A.; Elgorban, A.M.; Anandan, S.; Ashokkumar, M. Mercury removal from aqueous solution using petal-like MoS2 nanosheets. Front. Environ. Sci. Eng. 2021, 15, 1–10. [Google Scholar] [CrossRef]
  74. Zeng, M.; Yang, B.; Yan, H.; Qu, H.; Hu, Y. Efficient recovery of Ag(I) from aqueous solution using MoS2 nanosheets: Adsorption study and DFT calculation. Chem. Phys. Lett. 2020, 757, 137865. [Google Scholar] [CrossRef]
  75. Wang, Z.; Zhu, W.; Qiu, Y.; Yi, X.; Bussche, A.; Kane, A.; Gao, H.; Koski, K.; Hurt, R. Biological and environmental interactions of emerging two-dimensional nanomaterials. Chem. Soc. Rev. 2016, 45, 1750–1780. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Wu, P.R.; Liu, Z.; Cheng, Z.L. A top-down exfoliation for MoS2 nanosheets based on Li+/Na+ -intercalated and shearing synergistic process. Mater. Lett. 2019, 248, 236–240. [Google Scholar] [CrossRef]
  77. Deepak, D.R.; Nair, M.G.; Halder, S.; Sharma, A.L.; Mohapatra, S.R. Liquid phase exfoliation of MoS2 nano-sheets and observation of resistive switching memory in MoS2 Nano-sheets-PVDF-HFP composite films. Mater. Today Proc. 2019, 18, 5447–5453. [Google Scholar] [CrossRef]
  78. Hai, N.Q.; Kwon, S.H.; Kim, H.; Kim, I.T.; Lee, S.G.; Hur, J. High-performance MoS2-based nanocomposite anode prepared by high-energy mechanical milling: The effect of carbonaceous matrix on MoS2. Electrochim. Acta. 2018, 260, 129–138. [Google Scholar] [CrossRef]
  79. Li, Y.; Yin, X.; Huang, X.; Liu, X.; Wu, W. Efficient and scalable preparation of MoS2 nanosheet/carbon nanotube composites for hydrogen evolution reaction. Int. J. Hydrog. Energy 2020, 45, 16489–16499. [Google Scholar] [CrossRef]
  80. Yang, X.; Chen, Z.; Fang, J.; Yang, Q.; Zhao, W.; Qian, X.; Liu, C.; Zhou, D.; Tao, S.; Liu, X. Efficient exfoliation to MoS2 nanosheets by salt-assisted refluxing and ultrasonication with photocatalytic application. Mater. Lett. 2019, 255, 126596. [Google Scholar] [CrossRef]
  81. Tsigkourakos, M.; Kainourgiaki, M.; Skotadis, E.; Giannakopoulos, K.P.; Tsoukalas, D.; Raptis, Y.S. Capping technique for chemical vapor deposition of large and uniform MoS2 flakes. Thin Solid Film 2021, 733, 138808. [Google Scholar] [CrossRef]
  82. Luo, S.; Cullen, C.P.; Guo, G.; Zhong, J.; Duesberg, G.S. Investigation of growth-induced strain in monolayer MoS2 grown by chemical vapor deposition. Appl. Surf. Sci. 2020, 508, 145126. [Google Scholar] [CrossRef] [Green Version]
  83. Luo, L.; Shi, M.; Zhao, S.; Tan, W.; Lin, X.; Wang, H.; Jiang, F. Hydrothermal synthesis of MoS2 with controllable morphologies and its adsorption properties for bisphenol A. J. Saudi Chem. Soc. 2019, 23, 762–773. [Google Scholar] [CrossRef]
  84. Shan, J.; Li, J.; Chu, X.; Xu, M.; Jin, F.; Wang, X.; Ma, L.; Fang, X.; Wei, Z.; Wang, X. High sensitivity glucose detection at extremely low concentrations using a MoS2-based field-effect transistor. RSC Adv. 2018, 8, 7942–7948. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Ambrosi, A.; Chia, X.; Sofer, Z.; Pumera, M. Enhancement of electrochemical and catalytic properties of MoS2 through ball-milling. Electrochem. Commun. 2015, 54, 36–40. [Google Scholar] [CrossRef]
  86. Tayyebi, A.; Ogino, N.; Hayashi, T.; Komatsu, N. Size-controlled MoS2 nanosheet through ball milling exfoliation: Parameter optimization, structural characterization and electrocatalytic application. Nanotechnology 2020, 31, 75704. [Google Scholar] [CrossRef]
  87. Ma, H.; Ben, S.; Shen, Z.; Zhang, X.; Wu, C.; Liao, S.; An, F. Investigating the exfoliation behavior of MoS2 and graphite in water: A comparative study. Appl. Surf. Sci. 2020, 512, 145588. [Google Scholar] [CrossRef]
  88. Manamel, L.T.; Mukherjee, A.; Das, B.C. Two-dimensional nanohybrid of MoS2 and Rose Bengal: Facile solution growth and band structure probing. Appl. Surf. Sci. 2020, 530, 147063. [Google Scholar] [CrossRef]
  89. Singla, R.; Kumar, S.; Hackett, T.A.; Reshak, A.H.; Kashyap, M.K. Genesis of magnetism in graphene/MoS2 van der Waals heterostructures via interface engineering using Cr-adsorption. J. Alloys Compd. 2021, 859, 157776. [Google Scholar] [CrossRef]
  90. Li, H.; Zhu, X.; Tang, Z.K.; Zhang, X.H. Low-temperature photoluminescence emission of monolayer MoS2 on diverse substrates grown by CVD. J. Lumin. 2018, 199, 210–215. [Google Scholar] [CrossRef]
  91. Park, H.J.; Kim, M.S.; Kim, J.; Joo, J. Photo-responsive transistors of CVD grown single-layer MoS2 and its nanoscale optical characteristics. Curr. Appl. Physics 2016, 16, 1320–1325. [Google Scholar] [CrossRef]
  92. Khabiri, G.; Aboraia, A.M.; Soliman, M.; Guda, A.A.; Butova, V.V.; Yahia, I.S.; Soldatov, A.V. A novel α-Fe2O3@MoS2QDs heterostructure for enhanced visible-light photocatalytic performance using ultrasonication approach. Ceram. Int. 2020, 46, 19600–19608. [Google Scholar] [CrossRef]
  93. Gang, R.; Xu, L.; Xia, Y.; Cai, J.; Zhan, L.; Wang, S.; Li, R. Fabrication of MoS2 QDs/ZnO nanosheet 0D/2D heterojunction photocatalysts for organic dyes and gaseous heavy metal removal. J. Colloid. Interface Sci. 2020, 579, 853–861. [Google Scholar] [CrossRef]
  94. Wang, H.; Zhou, Y.; Xing, H.; Yang, X.; Zong, Y.; Feng, J.; Zhu, X.; Shi, Z.; Li, X.; Zheng, X. Construction of flower-like core-shell Fe3O4@2H-MoS2 heterostructures: Boosting the interfacial polarization for high-performance microwave absorption. Ceram Int. 2022, 48, 9918–9926. [Google Scholar] [CrossRef]
  95. Wu, Y.; Su, M.; Chen, J.; Xu, Z.; Tang, J.; Chang, X.; Chen, D. Superior adsorption of methyl orange by h-MoS2 microspheres: Isotherm, kinetics, and thermodynamic studies. Dye. Pigment. 2019, 170, 1–8. [Google Scholar] [CrossRef]
  96. Liu, P.; Cai, S.; Zuo, Y.; Tian, M.; Wang, Z.; Ling, L.; Sun, X. Synthesis of interlayer expanded MoS2 by sulfurization of MoO3 with enhanced sodium-ion storage. J Alloys Compd. 2022, 895, 162691. [Google Scholar] [CrossRef]
  97. Ai, K.; Ruan, C.; Shen, M.; Lu, L. MoS2 nanosheets with widened interlayer spacing for high-efficiency removal of mercury in aquatic systems. Adv. Funct. Mater. 2016, 26, 5542–5549. [Google Scholar] [CrossRef]
  98. Yuan, W.; Kuang, J.; Yu, M.; Huang, Z.; Zou, Z.; Zhu, L. Facile preparation of MoS2@Kaolin composite by one-step hydrothermal method for efficient removal of Pb(II). J. Hazard. Mater. 2021, 405, 124261. [Google Scholar] [CrossRef]
  99. Mário, E.D.A.; Liu, C.; Ezugwu, C.I.; Mao, S.; Jia, F.; Song, S. Molybdenum disulfide/montmorillonite composite as a highly efficient adsorbent for mercury removal from wastewater. Appl. Clay. Sci. 2020, 184, 105370. [Google Scholar] [CrossRef]
  100. Song, H.J.; You, S.; Jia, X.H.; Yang, J. MoS2 nanosheets decorated with magnetic Fe3O4 nanoparticles and their ultrafast adsorption for wastewater treatment. Ceram. Int. 2015, 41, 13896–13902. [Google Scholar] [CrossRef]
  101. Chen, Y.; Song, B.; Tang, X.; Lu, L.; Xue, J. Ultrasmall Fe3O4 nanoparticle/MoS2 nanosheet composites with superior performances for lithium ion batteries. Small 2014, 10, 1536–1543. [Google Scholar] [CrossRef] [PubMed]
  102. Xin, X.; Song, Y.; Guo, S.; Zhang, Y.; Wang, B.; Wang, Y.; Li, X. One-step synthesis of P-doped MoS2 for efficient photocatalytic hydrogen production. J. Alloys. Compd. 2020, 829, 154635. [Google Scholar] [CrossRef]
  103. Liu, F.; Wang, N.; Shi, C.; Sha, J.; Ma, L.; Liu, E.; Zhao, N. Phosphorus doping of 3D structural MoS2 to promote catalytic activity for lithium-sulfur batteries. Chem. Eng. J. 2022, 431, 133923. [Google Scholar] [CrossRef]
  104. Li, J.; He, T.; Zhao, Y.; Zhang, X.; Zhong, W.; Zhang, X.; Ren, J.; Chen, Y. In-situ N-doped ultrathin MoS2 anchored on N-doped carbon nanotubes skeleton by Mo-N bonds for fast pseudocapacitive sodium storage. J Alloys Compd. 2022, 897, 163170. [Google Scholar] [CrossRef]
  105. Zhang, Z.; Dong, Y.; Liu, G.; Li, J.; Sun, H.; Luo, H.; Liu, S. The ultrafine monolayer 1 T/2H-MoS2: Preparation, characterization and amazing photocatalytic characteristics. Colloids Surf. A Phys. Eng. Asp. 2020, 589, 124431. [Google Scholar] [CrossRef]
  106. Luo, N.; Chen, C.; Yang, D.; Hu, W.; Dong, F. S defect-rich ultrathin 2D MoS2: The role of S point-defects and S stripping-defects in the removal of Cr(VI) via synergistic adsorption and photocatalysis. Appl. Catal. B. 2021, 299, 120664. [Google Scholar] [CrossRef]
  107. Liu, S.; Nie, C.; Zhou, D.; Shen, J.; Feng, S. Direct growth of vertical structure MoS2 nanosheets array film via CVD method for photodetection. Physica E Low Dimens. Syst. Nanostruct. 2020, 117, 113592. [Google Scholar] [CrossRef]
  108. Cui, B.; Cai, X.; Wang, W.; Saha, P.; Wang, G. Nano storage-boxes constructed by the vertical growth of MoS2 on graphene for high-performance Li-S batteries. J. Energy Chem. 2022, 66, 91–99. [Google Scholar] [CrossRef]
  109. Zhi, L.; Zuo, W.; Chen, F.; Wang, B. 3D MoS2 composition aerogels as chemosensors and adsorbents for colorimetric detection and high-capacity adsorption of Hg2+. ACS Sustain. Chem. Eng. 2016, 4, 3398–3408. [Google Scholar] [CrossRef]
  110. Zhuang, Y.T.; Zhang, X.; Wang, D.H.; Yu, Y.L.; Wang, J.H. Three-dimensional molybdenum disulfide/graphene hydrogel with tunable heterointerfaces for high selective Hg(II) scavenging. J. Colloid. Interface Sci. 2018, 514, 715–722. [Google Scholar] [CrossRef]
  111. Wang, Z.; Zhang, J.; Wen, T.; Liu, X.; Wang, Y.; Yang, H.; Sun, J.; Feng, J.; Dong, S.; Sun, J. Highly effective remediation of Pb(II) and Hg(II) contaminated wastewater and soil by flower-like magnetic MoS2 nanohybrid. Sci. Total Environ. 2020, 699, 134341. [Google Scholar] [CrossRef] [PubMed]
  112. Chen, H.; Zhang, Z.; Zhong, X.; Zhuo, Z.; Tian, S.; Fu, S.; Chen, Y.; Liu, Y. Constructing MoS2/Lignin derived carbon nanocomposites for highly efficient removal of Cr(VI) from aqueous environment. J. Hazard. Mater. 2021, 408, 124847. [Google Scholar] [CrossRef] [PubMed]
  113. Gao, D.; Zhang, Y.; Yan, H.; Li, B.; He, Y.; Song, P.; Wang, R. Construction of UiO-66@MoS2 flowerlike hybrids through electrostatically induced self-assembly with enhanced photodegradation activity towards lomefloxacin. Sep. Purif. Technol. 2021, 265, 118486. [Google Scholar] [CrossRef]
  114. Shi, W.; Chu, Y.; Xia, M.; Wang, F.; Fu, C. The adsorption performance and micro-mechanism of MoS2/montmorillonite composite to atenolol and acebutolol: Adsorption experiments and a novel visual study of interaction. Ecotoxicol. Environ. Saf. 2021, 213, 111993. [Google Scholar] [CrossRef] [PubMed]
  115. Jayadharan Salini, A.N.; Ramachandran, A.; Sadasivakurup, S.; Yesodha, S.K. Versatile MoS2 hollow nanoroses for a quick-witted removal of Hg (II), Pb (II) and Ag (I) from water and the mechanism: Affinity or Electrochemistry? Appl. Mater. Today 2020, 20, 100642. [Google Scholar] [CrossRef]
  116. Prieto-Rodriguez, L.; Miralles-Cuevas, S.; Oller, I.; Agüera, A.; Puma, G.L.; Malato, S. Treatment of emerging contaminants in wastewater treatment plants (WWTP) effluents by solar photocatalysis using low TiO2 concentrations. J. Hazard. Mater. 2012, 211–212, 131–137. [Google Scholar] [CrossRef] [PubMed]
  117. Yuan, Y.J.; Lu, H.W.; Yu, Z.T.; Zou, Z.G. Noble-metal-free molybdenum disulfide cocatalyst for photocatalytic hydrogen production. Chem. Sus. Chem. 2015, 8, 4113–4127. [Google Scholar] [CrossRef]
  118. Chen, L.; Chuang, Y.; Nguyen, T.B.; Chang, J.H.; Lam, S.S.; Chen, C.W.; Dong, C.D. Novel molybdenum disulfide heterostructure nanohybrids with enhanced visible-light-induced photocatalytic activity towards organic dyes. J. Alloys. Compd. 2020, 848, 156448. [Google Scholar] [CrossRef]
  119. Abraham, T.; Priyanka, R.N.; Joseph, S.; Plathanam, N.J.; Gigimol, M.G.; Mathew, B. Flower-like MoS2/BiFeO3 doped silver orthophosphate catalyst for visible-light assisted treatment of refractory organic pollutants. Appl. Mater. Today 2020, 21, 100845. [Google Scholar] [CrossRef]
  120. Fu, Y.; Ren, Z.; Wu, J.; Li, Y.; Liu, W.; Li, P.; Xing, L.; Ma, J.; Wang, H.; Xue, X. Direct Z-scheme heterojunction of ZnO/MoS2 nanoarrays realized by flowing-induced piezoelectric field for enhanced sunlight photocatalytic performances. Appl. Catal. B Environ. 2021, 285, 119785. [Google Scholar] [CrossRef]
  121. Quan, Y.; Su, R.; Hu, M.; Lang, J.; Fan, H.; Shen, H.; Gao, M.; Li, B.; Liu, Y.; Yang, J. Construction of an MZO heterojunction system with improved photocatalytic activity for degradation of organic dyes. Cryst. Eng. Comm. 2020, 22, 7059–7065. [Google Scholar] [CrossRef]
  122. Guo, P.; Zhao, F.; Hu, X. Fabrication of a direct Z-scheme heterojunction between MoS2 and B/Eu-g- C3N4 for an enhanced photocatalytic performance toward tetracycline degradation. J. Alloys. Compd. 2021, 867, 159044. [Google Scholar] [CrossRef]
  123. Ahamad, T.; Naushad, M.; Alzaharani, Y.; Alshehri, S.M. Photocatalytic degradation of bisphenol-A with g-C3N4/MoS2-PANI nanocomposite: Kinetics, main active species, intermediates and pathways. J. Mol. Liq. 2020, 311, 113339. [Google Scholar] [CrossRef]
  124. Putritama, V.; Fauzia, V.; Supangat, A. The effect of the layer number of MoS2 nanosheets on the photocatalytic efficiency of ZnO/MoS2. Surf. Interfaces 2020, 21, 100745. [Google Scholar] [CrossRef]
  125. Liang, H.; Bai, J.; Xu, T.; Li, C. Enhancing photocatalytic performance of heterostructure MoS2/g-C3N4 embeded in PAN frameworks by electrospining process. Mater. Sci. Semicond. Process 2021, 121, 105414. [Google Scholar] [CrossRef]
  126. Zhang, X.; Xia, M.; Wang, F.; Lei, W. Cu2O/MoS2 composites: A novel photocatalyst for photocatalytic degradation of organic dyes under visible light. Ionics 2020, 26, 6359–6369. [Google Scholar] [CrossRef]
  127. Mahalakshmi, G.; Rajeswari, M.; Ponnarasi, P. Fabrication of dandelion clock-inspired preparation of core-shell TiO2@MoS2 composites for unprecedented high visible light-driven photocatalytic performance. J. Mater. Sci. Mater. Electron. 2020, 31, 22252–22264. [Google Scholar] [CrossRef]
  128. Mahalakshmi, G.; Rajeswari, M.; Ponnarasi, P. Synthesis of few-layer g-C3N4 nanosheets-coated MoS2/TiO2 heterojunction photocatalysts for photo-degradation of methyl orange (MO) and 4-nitrophenol (4-NP) pollutants. Inorg. Chem. Commun. 2020, 120, 108146. [Google Scholar] [CrossRef]
  129. Adhikari, S.; Mandai, S.; Kim, D.H. Free-standing Ag nanoparticle-decorated MoS2 microflowers grown on carbon cloth for photocatalytic oxidation of Rhodamine, B. Korean J. Chem. Eng. 2020, 37, 2359–2367. [Google Scholar] [CrossRef]
  130. Wang, Y.; Sun, S.; Liu, Y.; Zhang, Y.; Xia, J.; Yang, Q. TiO2 coupled to predominantly metallic MoS2 for photocatalytic degradation of rhodamine B. J. Mater. Sci. 2020, 55, 12274–12286. [Google Scholar] [CrossRef]
  131. Kumaresan, N.; Karuppasamy, P.; Pandian, M.S.; Ramasamy, P. Formation of face-contact interaction in 2D/2D/2D heterostructure ternary nanocomposites of g-C3N4/MoS2/GO for effective photocatalytic activity against the organic pollutants under the visible light irradiation. J. Mater. Sci. Mater. Electron. 2022, 33, 11970–11988. [Google Scholar] [CrossRef]
  132. Rao Akshatha, S.; Sreenivasa, S.; Parashuram, L.; Raghu, M.S.; Yogesh Kumar, K.; Madhu Chakrapani Rao, T. Visible-light-induced photochemical hydrogen evolution and degradation of crystal violet dye by interwoven layered MoS2/Wurtzite ZnS Heterostructure Photocatalyst. Chem. Select. 2020, 5, 6918–6926. [Google Scholar]
  133. Jaleel, U.C.J.R.; Devi, K.R.S.; Madhushree, R.; Pinheiro, D. Statistical and experimental studies of MoS2/g-C3N4/TiO2: A ternary Z-scheme hybrid composite. J. Mater. Sci. 2021, 56, 6922–6944. [Google Scholar] [CrossRef]
  134. Shi, Z.; Zhang, Y.; Duoerkun, G.; Cao, W.; Liu, T.; Zhang, L.; Liu, J.; Li, M.; Chen, Z. Fabrication of MoS2/BiOBr heterojunctions on carbon fibers as a weaveable photocatalyst for tetracycline hydrochloride 65 degradation and Cr (VI) reduction under visible light. Environ. Sci. Nano. 2020, 7, 2708–2722. [Google Scholar] [CrossRef]
  135. Guo, J.; Yang, C.; Sun, Z.; Yang, Z.; Wang, L.; Lu, C.; Ma, Z.; Guo, F. Ternary Fe3O4/MoS2/BiVO4 nanocomposites: Novel magnetically separable visible light-driven photocatalyst for efficiently degradation of antibiotic wastewater through p–n heterojunction. J. Mater. Sci. Mater. Electron. 2020, 31, 16746–16758. [Google Scholar] [CrossRef]
  136. Liu, J.; Lin, H.; He, Y.; Dong, Y.; Rose, E.; Menzembere, G.Y. Novel CoS2/MoS2@Zeolite with excellent adsorption and photocatalytic performance for tetracycline removal in simulated wastewater. J. Clean. Prod. 2020, 260, 121047. [Google Scholar] [CrossRef]
  137. Chen, L.; Huang, C.P.; Chuang, Y.; Nguyen, T.B.; Chen, C.W.; Dong, C.D. Z-Scheme MoS2/TiO2/graphene nanohybrid photocatalysts for visible light-induced degradation for highly efficient water disinfection and antibacterial activity. New J. Chem. 2022, 46, 14159–14169. [Google Scholar] [CrossRef]
  138. Chen, L.; Chuang, Y.; Chen, C.W.; Dong, C.D. Facile synthesis of MoS2/ZnO quantum dots for enhanced visible-light photocatalytic performance and antibacterial applications. Nano Struct. Nano Objects 2022, 30, 100873. [Google Scholar] [CrossRef]
  139. Vaizoğullar, A.İ. Ternary CdS/MoS2/ZnO photocatalyst: Synthesis, characterization and degradation of ofloxacin under visible light irradiation. J. Inorg. Organomet. Polym. Mater. 2020, 30, 4129–4141. [Google Scholar] [CrossRef]
  140. Zou, X.; Zhao, X.; Zhang, J.; Lv, W.; Qiu, L.; Zhang, Z. Photocatalytic degradation of ranitidine and reduction of nitrosamine dimethylamine formation potential over MXene–Ti3C2/MoS2 under visible light irradiation. J. Hazard. Mater. 2021, 413, 125424. [Google Scholar] [CrossRef]
  141. Ghasemipour, P.; Fattahi, M.; Rasekh, B.; Yazdian, F. Developing the ternary ZnO doped MoS2 nanostructures grafted on CNT and reduced graphene oxide (RGO) for photocatalytic degradation of aniline. Sci. Rep. 2020, 10, 4414. [Google Scholar] [CrossRef] [Green Version]
  142. Wang, Y.; Tang, X.; Liu, Z.; Yan, Y.; Yang, B.; Zhu, Z. Fabrication of a Z-scheme MoS2/CuO heterojunction for enhanced 2-mercaptobenzothiazole degradation activity and mechanism insight. New J. Chem. 2020, 44, 18264–18273. [Google Scholar] [CrossRef]
  143. Sajjad, M.; Tahir, M.B.; Mubeen, I.; Kabli, M.R. Tailorable and rationally designed MoS2 based heterostructure photocatalyst for efficient photocatalytic degradation of phenol under the visible light. J. Inorg. Organomet. Polym. Mater. 2020, 30, 3965–3972. [Google Scholar] [CrossRef]
  144. Bhati, A.; Anand, S.R.; Saini, D.; Gunture, S.; Sonkar, K. Sunlight-induced photoreduction of Cr(VI) to Cr(III) in wastewater by nitrogen-phosphorus-doped carbon dots. NPJ Clean Water 2019, 2, 12. [Google Scholar] [CrossRef]
  145. Kumari, A.; Chaudhary, D.R. Engineered microbes and evolving plastic bioremediation technology. INC. 2020, 417–443. [Google Scholar]
  146. Somasundar, Y.; Burton, A.E.; Mills, M.R.; Zhang, D.Z.; Ryabov, A.D.; Collins, T.J. Quantifying evolving toxicity in the TAML/peroxide mineralization of propranolol. iScience 2021, 24, 101897. [Google Scholar] [CrossRef]
  147. Zeng, Z.; Ye, S.; Wu, H.; Xiao, R.; Zeng, G.; Liang, J.; Zhang, C.; Yu, J.; Fang, Y.; Song, B. Research on the sustainable efficacy of g-MoS2 decorated biochar nanocomposites for removing tetracycline hydrochloride from antibiotic-polluted aqueous solution. Sci. Total Environ. 2019, 648, 206–217. [Google Scholar] [CrossRef]
  148. Huang, S.; Chen, C.; Tsai, H.; Shaya, J.; Lu, C. Photocatalytic degradation of thiobencarb by a visible light-driven MoS2 photocatalyst. Sep. Purif. Technol. 2018, 197, 147–155. [Google Scholar] [CrossRef]
  149. Chandrabose, G.; Dey, A.; Gaur, S.S.; Pitchaimuthu, S.; Jagadeesan, H.; Braithwaite, N.J.; Selvaraj, V.; Kumar, V.; Krishnamurthy, S. Removal and degradation of mixed dye pollutants by integrated adsorption photocatalysis technique using 2-D MoS2/TiO2 nanocomposite. Chemosphere 2021, 279, 130467. [Google Scholar] [CrossRef]
  150. Saleh, T.A. Trends in the sample preparation and analysis of nanomaterials as environmental contaminants. Trends Environ. Anal. Chem. 2020, 28, e00101. [Google Scholar] [CrossRef]
  151. Louie, S.M.; Ma, R.; Lowry, G.V. Transformations of nanomaterials in the environment. Front. Nanosci. 2014, 7, 55–87. [Google Scholar]
  152. Cicuéndez, M.; Silva, V.S.; Santos, J.; Coimbra, A.; Oliveira, H.; Ayán-Varela, M.; Paredes, J.I.; VillarRodil, S.; Vila, M. MoS2 flakes stabilized with DNA/RNA nucleotides: In vitro cell response. Mater. Sci. Eng. C 2019, 100, 11–22. [Google Scholar] [CrossRef]
Figure 1. Atomic structures of various polymorphs of MoS2. (a) Monolayer structures; (b) bilayer and trilayer structures (yellow—S atoms; purple—Mo atoms). A, A’, B and C represent S-Mo-S layers. AA’, AB and ABC represent the stacking order sequence Reproduced from [44], with permission from John Wiley & Sons, Copyright 2018 Wiley–VCH GmbH.
Figure 1. Atomic structures of various polymorphs of MoS2. (a) Monolayer structures; (b) bilayer and trilayer structures (yellow—S atoms; purple—Mo atoms). A, A’, B and C represent S-Mo-S layers. AA’, AB and ABC represent the stacking order sequence Reproduced from [44], with permission from John Wiley & Sons, Copyright 2018 Wiley–VCH GmbH.
Molecules 27 06782 g001
Figure 2. Electron microscopy images of MoS2, synthesized using different methods: (a) Ball-milling method. Reprinted from [78]. Copyright (2018), with permission from Elsevier. (b) Ball-milling method. Reprinted from [79]. Copyright (2020), with permission from Elsevier. (c) Liquid phase exfoliation method. Reprinted from [80]. Copyright (2019), with permission from Elsevier. (d) Chemical vapor deposition method. Reprinted from [81]. Copyright (2021), with permission from Elsevier. (e) Chemical vapor deposition method. Adapted from [82]. Copyright (2020), with permission from Elsevier. (f) Hydrothermal method. Reprinted from [83]. Copyright (2019), with permission from Elsevier.
Figure 2. Electron microscopy images of MoS2, synthesized using different methods: (a) Ball-milling method. Reprinted from [78]. Copyright (2018), with permission from Elsevier. (b) Ball-milling method. Reprinted from [79]. Copyright (2020), with permission from Elsevier. (c) Liquid phase exfoliation method. Reprinted from [80]. Copyright (2019), with permission from Elsevier. (d) Chemical vapor deposition method. Reprinted from [81]. Copyright (2021), with permission from Elsevier. (e) Chemical vapor deposition method. Adapted from [82]. Copyright (2020), with permission from Elsevier. (f) Hydrothermal method. Reprinted from [83]. Copyright (2019), with permission from Elsevier.
Molecules 27 06782 g002
Figure 3. SEM images of MoS2 composites synthesized via the hydrothermal method. (a) S/Mo molar ratio 1:1; (b) S/Mo molar ratio 4:1; (c) synthesis time: 25 h at 150 °C; (d) synthesis time: 37 h at 180 °C; (e) synthesis time: 47 h at 240 °C. Reprinted from [83]. Copyright (2019), with permission from Elsevier.
Figure 3. SEM images of MoS2 composites synthesized via the hydrothermal method. (a) S/Mo molar ratio 1:1; (b) S/Mo molar ratio 4:1; (c) synthesis time: 25 h at 150 °C; (d) synthesis time: 37 h at 180 °C; (e) synthesis time: 47 h at 240 °C. Reprinted from [83]. Copyright (2019), with permission from Elsevier.
Molecules 27 06782 g003
Figure 4. Schematics of the proposed photocatalysis mechanism of α-Fe2O3/MoS2QDs. Reproduced from [92], with permission from John Wiley & Sons. Copyright 2018 Wiley-VCH GmbH.
Figure 4. Schematics of the proposed photocatalysis mechanism of α-Fe2O3/MoS2QDs. Reproduced from [92], with permission from John Wiley & Sons. Copyright 2018 Wiley-VCH GmbH.
Molecules 27 06782 g004
Table 1. Kinetics and equilibrium parameters from the fitted models in adsorption studies for the removal of metal ion species and organic contaminants, using MoS2-based composites.
Table 1. Kinetics and equilibrium parameters from the fitted models in adsorption studies for the removal of metal ion species and organic contaminants, using MoS2-based composites.
CompositeMoS2 Preparation MethodAdsorption CompoundTemperature
(°C)
Adsorption KineticsAdsorption IsothermsRef.
PFOPSOLangmuirFreundlich
Metal species
MoS2HTHg(II)27qe (mg/g)208.5303 [73]
k1 (min−1)0.032-
k2 (g/(mg min))-2.710
R20.9410.998
Au/Fe3O4/MoS2CAs CommercialHg(II) qe (mg/g)0.47920.408qm (mg/g)1527n3.040[109]
k1 (min−1)0.089-KL (L/mg)0.083KF (mg/g)200.85
k2 (g/(mg min))-1.2R20.999R20.774
R20.7790.999
3D MoS2-rGO HTHg(II)25qe (mg/g) 30.55qm (mg/g)400n1.046[110]
k1 (min−1)0.117-KL (L/mg)65.68KF (mg/g)5.936
k2 (g/(mg min))-3.75 × 10−3R20.9998R20.999
R20.7850.999
W-DR-N-MoS2HTHg(II) qe (mg/g)--qm (mg/g)2562.8n [97]
k1 (min−1)--KL (L/mg)3.029KF (mg/g)
k2 (g/(mg min))-1.42R20.999R2
R2-0.9999
MoS2/Fe3O4HTHg(II)25qe (mg/g)84.46254.3qm (mg/g)428.9N198.8[111]
k1 (min−1)0.142-KL (L/mg)1.586KF (mg/g)5.085
k2 (g/(mg min))-3.9 × 10−3R21.000R20.723
R20.9221.000
MoS2/MMT HTHg(II)35qe (mg/g)1119.91390.81qm (mg/g)2055.9n0.759[99]
k1 (min−1)5.426 × 10−3-KL (L/mg)0.030KF (mg/g)3.204
k2 (g/(mg min))-3.840 × 10−6R20.944R20.838
R20.9820.96089
MoS2-clinoptilolite CommercialPb(II)25qe (mg/g)1.31750qm (mg/g)3.45n2.048[8]
k1 (min−1)6.17 × 10−2-KL (L/mg)0.38KF (mg/g)1.527
k2 (g/(mg min))-2.0 × 10−2R20.964R20.837
R20.7231.000
MoS2/Fe3O4HTPb(II)25qe (mg/g)119.1199.3qm (mg/g)263.6n61.62[111]
k1 (min−1)0.011-KL (L/mg)0.1468KF (mg/g)3.354
k2 (g/(mg min))-1.7 × 10−4R21.000R20.841
R20.9190.922
MoS2-kaolin HTPb(II)25qe (mg/g)12.2365.94qm (mg/g)280.39n5.71[98]
k1 (min−1)−0.086 KL (L/mg)2.93KF (mg/g)105.43
k2 (g/(mg min)) 0.015R20.863R20.790
R20.04840.9922
MoS2/LDC HTCr(IV)25qe (mg/g)7.65204.08qm (mg/g)306.85n8.333[112]
k1 (min−1)1.4 × 10−2-KL (L/mg)0.97KF (mg1-n Ln/g)20.73
k2 (g/(mg min))-4 × 10−4R20.86R20.76
R20.680.999
2D MoS2HTAg(I)25qe (mg/g)392.73410.24qm (mg/g)813.01n3.62[74]
k1 (min−1)0.921-KL (L/mg)0.12KF (mg/g)206.51
k2 (g/(mg min))-3.0 × 10−3R20.999R20.911
R20.9880.997
chitosan-coated MoS2 1:2 Precipitation methodAu(III)30qe (mg/g)1230.841338.55qm (mg/g)3434.96n807.74[28]
k1 (min−1)0.074-KL (L/mg)0.043KF (mg/g)3.857
k2 (g/(mg min))-6.45 × 10−5R20.991R20.938
R20.7240.901
Organic Contaminants
MoS2HTBisphenol A25qe (mg/g)28.5730.25qm (mg/g) n [83]
k1 (min−1)--KL (L/mg) KF (mg/g)
k2 (g/(mg min))--R2 R2
R20.9010.998
rGO/MoS2HTOfloxacin qe (mg/g)-43.104qm (mg/g)37.31n3.84[7]
k1 (min−1)--KL (L/mg)0.46KF (mg/g)15.07
k2 (g/(mg min))-2.69 × 10−4R20.989R20.845
R2-0.918
UiO−66/MoS2HTLomefloxanin qe (mg/g)45.680.67qm (mg/g)32.36n5.152[113]
k1 (min−1)0.2434-KL (L/mg)0.28606KF (mg/(g(L/mg)1/n))14.957
k2 (g/(mg min))-1.7 × 10−4R20.786R20.940
R20.9740.973
MoS2/MTT HTAtenolol25qe (mg/g)88.8596.62qm (mg/g)145.6n3.363[114]
k1 (min−1)0.151-KL (L/mg)0.197KF (L/g)41.38
k2 (g/(mg min))-2.364 × 10−3R20.974R20.859
R20.9570.991
MoS2/MTT HTAcebutolol25qe (mg/g)63.8569.25qm (mg/g)130.8n2.772[114]
k1 (min−1)0.287-KL (L/mg)0.107KF (L/g)25.42
k2 (g/(mg min))-5.617 × 10−3R20.945R20.831
R20.8290.982
h-MoS2HTMO25qe (mg/g)12.8642.44qm (mg/g)38.11n1.897[91]
k1 (min−1)0.518-KL (L/mg)0.285KF (mg/(g(L/mg)1/n))5.342
k2 (g/(mg min))-9.35 × 10−3R20.952R20.992
R20.8480.999
h-MoS2—hollow MoS2 microspheres; HT—hydrothermal method; k1—rate constant of pseudo-first-order kinetic model; k2—rate constant of pseudo-second-order kinetic model; KF–Freundlich constant; KL—Langmuir constant; LDC—lignin-derived carbon; MMT—montmorillonite; MO–methyl orange; MoS2CAs—MoS2 composite aerogel; n–heterogeneity factor; PFO—pseudo-first-order kinetic model; PSO—pseudo-second-order kinetic model; qe—adsorption capacity at equilibrium; qm—maximum adsorption capacity; rGO—reduced graphene oxide; W-DR-N- MoS2—widened defect-rich nano MoS2.
Table 2. Characteristics of MoS2-based catalysts and the respective photocatalytic performance (in optimized conditions).
Table 2. Characteristics of MoS2-based catalysts and the respective photocatalytic performance (in optimized conditions).
MoS2-Based CatalystContaminantMoS2 Production MethodBand Gap (eV)CSA (m2/g)IrradiationIrradiation Time (min)Removal RatePFO Kinetics
k (min−1)
Ref.
Dyes
α-Fe2O3/MoS2QDs MBHT2.225.852Vis (400–700 nm)1
120
83%
91%
1st min 0.892
> 28 min 0.0049
[92]
MoS2/ZnO MBHT-14.785Vis4097%0.0701[121]
5 wt% MoS2/BiFeO3/Ag3PO4MBC2.077.705UV-Vis597%0.625[119]
ZnO/MoS2 8 h MBLPE1.77-UV-Vis37.5Vis-85%
UV-78%
-[124]
Ni foam with ZnO/MoS2MBHT-4.85UV-Vis5072%-[120]
1.9% MoS2/g-C3N4/PAN MBHT2.7615.6Vis (>420 nm)12085%-[125]
MoS2/Fe3O4 (MF-17) MBHT-4,5Vis12098%-[25]
CM12 (Cu2O/MoS2) MBHTCu2O—2.15
MoS2—1.76
-Vis (>420 nm)3090%0.084[126]
MoS2-TiO2MOHT2.76103.5-6086%0.016[127]
CoO/meso-CN/MoS2 (1%) (S6) MOHT2.82–2.8637.64VIS6084%-[118]
5 wt% MoS2/BiFeO3/Ag3PO4MOC2.077.704UV-Vis598%0.376[119]
g-C3N4/MoS2/TiO2 (CMT5) MOHT2.6497.5Vis (400–700 nm)6098%0.061[128]
Ni foam with ZnO/MoS2MOHT-4.85UV-vis1093%0.059[122]
5% MoS2/ZnO RhBHT3.18-UV5095%0.057[119]
5 wt% MoS2/BiFeO3/Ag3PO4RhBC2.077.704UV-Vis597%0.676[119]
Ag/MoS2/CC RhBHT-27Vis20
40
90%
99%
72.1 × 10−3[129]
CTAB-MoS2-P25 RhBHT2.06--12091%-[130]
Ni foam with ZnO/MoS2RhBHT-4.85UV-vis808%-[120]
1.9% MoS2/g-C3N4/PAN RhBHT2.7615.6Vis (>420 nm)12048%0.006[125]
BiOI/MoS2 (BMS-8) RhBHTBiOI—1.42
MoS2—1.73
30.76Vis (>420 nm)90100%0.023[3]
MoS2/Fe3O4 (MF-17) RhBHT-4,5Vis12096%-[25]
g-C3N4/MoS2/GO (AT3G15) RhBHT2.05-Vis6099%-[131]
5wt% MoS2/BiFeO3/Ag3PO4Acrid red 18C2.077.7045UV-Vis798%0.484[119]
CoO/meso-CN/MoS2(1%) (S6) Methyl redHT2.82–2.8637.64Vis6096%0.072[118]
CoO/meso-CN/MoS2(1%) (S6) Congo redHT2.82–2.8637.64Vis6095%-[118]
g-C3N4/MoS2/TiO2 (CMT5) 4-NitrophenolHT2.6497.5Vis (400–700 nm))6087%-[128]
MoS2-ZnS Crystal violetHT--Vis4099%41.26 × 10−3[132]
Ni foam with ZnO/MoS2Congo redHT-4.85UV-Vis8077% [120]
MoS2/g-C3N4/TiO2Malachite greenHT2.42-Vis6086%0.045[133]
Antibiotics
MoS2/BiOBr/CF TCHTMoS2—1.81
BiOBr—2.88
-Vis12089%-[134]
Fe3O4/MoS2/BiVO4 (FMB3) TCHTMoS2—1.56
BiVO4—2.44
-Vis (>420 nm)12086%0.01576[135]
5 wt% MoS2/BiFeO3/Ag3PO4TCC2.077.704UV-Vis12090%-[119]
MoS2/Z-50 TCHT-17.32Vis (>420 nm)12097%-[136]
20 wt% MoS2/BEuCN TCHT2.8544.12Vis5099%0.087[122]
MoS2/TiO2/graphene (MTG-48) TCHT3.1758.6UV-Vis (300–750 nm)6092%0.05[137]
MoS2/ZnO QDs TCHT--Vis8096%0.01[138]
CdS/MoS2/ZnO(CMZ) Ofloxacinchemical co-precipitation methodCdS—2.145
MoS2—2.015
ZnO—2.981
-Vis (>420 nm)9089%0.024[139]
UiO-66/MoS2 (UMS-0.15) LomefloxacinHT-37.176Vis9087%-[113]
Other organic compounds
MXene-Ti3C2/MoS2 (MT-4) RanitidineHT1.5911.93Vis6088%-[140]
rGO10%/ZnO20%/MoS2AnilineHT2.24-Vis120100%-[141]
MoS2-TiO24-NitrophenolHT2.76103.5Vis6097%0.024[127]
MoS2/CuO-25% 2-MBTHTMoS2—1.52
CuO—2.16
47Vis12096%-[142]
Cu2S-1.0%MoS2PhenolC1.42-Vis9090%-[143]
g-C3N4/MoS2/GO (AT3G15) 4-CPHT2.05-Vis6089%-[131]
5 wt% MoS2/BiFeO3/Ag3PO42,4-DC2.077.704UV-Vis18090%-[119]
5 wt% MoS2/BiFeO3/Ag3PO4AcephateC2.077.704UV-Vis6085%-[119]
g-C3N4/MoS2−PANI Bisphenol-AHT2.67184.21Vis6093%0.040[123]
Metal ion species
MoS2/BiOBr/CF Cr(VI)HTMoS2—1.81
BiOBr—2.88
-Vis12085%-[134]
Ni foam with ZnO/MoS2Cr(VI)HT-4.85UV-Vis2575%-[120]
MoS2/TiO2/graphene (MTG-48) Pb(II)HT3.1758.6UV-Vis (300–750 nm)6088%-[137]
Note: 2,4-D—2,4-dichlorophenoxyacetic acid; 2-MBT—2-mercaptobenzothiazole; 4-CP—4-chlorophenol; C—commercial; CC—carbon cloth; CF—carbon fiber; CTAB- cetyltrimethyl ammonium bromide; CSA—catalyst-specific area; HT—hydrothermal method; LPE—liquid-phase exfoliation method; MB—methylene blue; MO—methyl orange; PAN—polyacrylonitrile; PANI—polyaniline; PFO—pseudo-first order; QDs—quantum dots; RhB—rhodamine B; Z-50—zeolite; TC—tetracycline.
Table 3. Photodegradation rate and mineralization evaluation on photocatalytic studies, with MoS2-based catalysts.
Table 3. Photodegradation rate and mineralization evaluation on photocatalytic studies, with MoS2-based catalysts.
MoS2-Based CatalystContaminantIrradiation Time (min)Photodegradation/Removal RateMineralization
Evaluation
Ref.
5 wt % MoS2/BiFeO3/Ag3PO4MB597%>85%[119]
CoO/meso-CN/MoS2(1%) (S6)MO6084%54%[118]
5 wt % MoS2/BiFeO3/Ag3PO4MO598%> 85%[119]
g-C3N4/MoS2/TiO2 (CMT5)MO6098%91%[128]
5 wt % MoS2/BiFeO3/Ag3PO4RhB597%>85%[119]
BiOI/MoS2 (BMS-8)RhB90100%78% *[3]
g-C3N4/MoS2/GO (AT3G15)RhB6099%85%[131]
5 wt % MoS2/BiFeO3/Ag3PO4Acrid red 18798%>85%[119]
CoO/meso-CN/MoS2(1%) (S6)Methyl red6096%80%[118]
CoO/meso-CN/MoS2(1%) (S6)Congo red6095%71%[118]
MoS2/BiOBr/CFTC12092%55% **[134]
5 wt % MoS2/BiFeO3/Ag3PO4TC12090%91%[119]
MoS2/TiO2/graphene (MTG-48)TC6092%33.8%[137]
MXene-Ti3C2/MoS2 (MT-4)Ranitidine6088%74%[140]
5 wt % MoS2/BiFeO3/Ag3PO42,4-D18090%87%[119]
5 wt % MoS2/BiFeO3/Ag3PO4Acephate6085%93%[119]
Note: 2,4-D—2,4-dichlorophenoxyacetic acid; CF—carbon fiber; MB—methylene blue; MO—methyl orange; RhB—rhodamine B; TC—tetracycline; * TOC after 60 min of irradiation; ** TOC after 7 h of irradiation.
Table 4. Adsorption and photocatalytic performance of MoS2-based nanocomposites and a comparison with bare MoS2.
Table 4. Adsorption and photocatalytic performance of MoS2-based nanocomposites and a comparison with bare MoS2.
Adsorption
CompositeContaminantSpecific Area (m2/g)Contact Time (min)Comparative Adsorption ParametersRef.
W-DR-N-MoS2Hg(II)--Interlaying spacing9.42 Å Maximum adsorption capacity (qmax) (mg g−1)2562.8[97]
MoS2Hg(II)--6.15 Å35.5
MoS2/MTTHg(II)--Adsorption capacity in equilibrium, according to the pseudo-first-order kinetic model (mg g−1)1119.94[99]
MoS2Hg(II)--936.62
MoS2-kaolinPb(II)14.5610Adsorption capacity
(mg g−1)
55.10 Removal rate89%[98]
MoS2Pb(II)13.081035.68 54%
MoS2/MTTAtenolol-150Adsorption capacity (mg g−1)132.08[114]
MoS2Atenolol-15074.23
MoS2/MTTAcebutolol-150Adsorption capacity (mg g−1)113.82[114]
MoS2Acebutolol-15036.05
Photocatalysis
CatalystContaminantBand Gap (eV)Specific Area (m2/g)Irradiation Time (min)Removal Rate (%)PFO Kinetics
k (min−1)
Ref.
MZO (MoS2/ZnO)MB-14.78540970.070[121]
MoS2MB-3.79540440.010
MoS2/Fe3O4 (MF-17)MB-4.512098-[25]
MoS2MB-9.012092-
MoS2-TiO2MO2.76103.560860.016[127]
MoS2MO3.2688.560530.008
BiOI/MoS2 (BMS-8)RhBBiOI 1.42
MoS2 1.73
30.76901000.023[3]
MoS2RhBMoS2 1.7349.5990600.009
MoS2/Fe3O4 (MF-17)RhB-4.512096-[25]
MoS2RhB-9.012082-
MoS2/TiO2/graphene (MTG-48)TC3.1758.660920.05[137]
MoS2TC2.6629.360540.018
MoS2-ZnSCrystal violet--4098.541.26 × 10−3[132]
MoS2Crystal violet--406024.23 × 10−3
MoS2-TiO24-Nitrophenol2.76103.560970.024[127]
MoS24-Nitrophenol3.2688.560590.009
UiO-66/MoS2 (UMS-0.15)Lomefloxacin-37.176908737.176[113]
MoS2Lomefloxacin-7.7759038.27.775
MoS2/CuO-25%2-MBTMoS2 1.52
CuO 2.16
4712096-[142]
MoS22-MBT1.5214.7612022-
Note: 2-MBT—2,4-D—2,4-dichlorophenoxyacetic acid; MB—methylene blue; MO—methyl orange; MMT—montmorillonite; RhB—rhodamine B; TC—tetracycline; W-DR-N-MoS2—Widened Defect Rich Nano MoS2.
Table 5. Reuse of MoS2-based sorbents and photocatalysts.
Table 5. Reuse of MoS2-based sorbents and photocatalysts.
Adsorption
MoS2-Based CompositeContaminantContact Time (min)Removal
1st Cycle (%)
Number of CyclesRemoval after Last Cycle (%)Ref.
Au/Fe3O4/MoS2CAsHg(II)3095–10010>95[109]
MoS2/Fe3O4Hg(II)-100–955identical[111]
MoS2-kaolinHg(II)5099577[98]
MoS2-rGOHg(II)7100685–90[110]
MoS2/MTTHg(II)-100485.2[99]
MoS2/Fe3O4Pb(II)-100–955identical[111]
MoS2/LDCCr(VI)3099490[112]
chitosan-coated MoS2 1:2Au(III)-98.9486.4[28]
MoS2-rGOOfloxacin-9284[7]
Photocatalysis
MoS2-Based CatalystContaminantIrradiation Time (min)Removal
1st Cycle (%)
Number of CyclesRemoval after Last Cycle (%)Ref.
α-Fe2O3/MoS2QDsMB1
120
83
91
31min 65
120 min 75
[92]
MZO (MoS2/ZnO)MB4097589.10[121]
ZnO/MoS2 8 hMB37.585565[124]
CM12 (Cu2O/MoS2)MO3090540[126]
CoO/meso-CN/MoS2(1%) (S6)MO6084580[118]
Ni foam with ZnO/MoS2MO 1092.7486.9[120]
5% MoS2/ZnORhB5095587[93]
CTAB-MoS2-P25RhB12091476[130]
1.9% MoS2/g-C3N4/PANRhB 12048443[125]
BiOI/MoS2 (BMS-8)RhB901006>90[3]
g-C3N4/MoS2/GO (AT3G15)RhB6099592[131]
CoO/meso-CN/MoS2(1%) (S6)Methyl red6096590[118]
CoO/meso-CN/MoS2(1%) (S6)Congo red6095588[118]
MoS2-ZnSCrystal violet4098.5490.5[132]
MoS2/g-C3N4/TiO2Malachite green6086470[133]
MoS2/BiOBr/CFTC 12089.0480.7[134]
Fe3O4/MoS2/BiVO4 (FMB3)TC12086.15>80[135]
MoS2/Z-50TC12096.8587.2[136]
20 wt% MS/BEuCNTC50993identical[120]
MoS2/TiO2/graphene (MTG-48)TC6092585[137]
MoS2/ZnO QDsTC8096589[138]
CdS/MoS2/ZnO(CMZ)Ofloxacin9089470–75[139]
UiO-66/MoS2 (UMS-0.15)Lomefloxacin9087479[113]
MXene-Ti3C2/MoS2 (MT-4)Ranitidine6088.4 80[140]
RGO10%/ZnO20%/MoS2Aniline1201005100[141]
MoS2/CuO-25%2-MBT12096591[142]
MoS2/BiOBr/CFCr(VI)12084.7476.6[134]
MoS2/TiO2/graphene (MTG-48)Pb(II)6088680[137]
Note: 2-MBT—2-mercaptobenzothiazole; CF—carbon fiber; CTAB—cetyltrimethyl ammonium bromide; LDC—lignin-derived carbon; MB—methylene blue; MO—methyl orange; MTT—montmorillonite; QDs—quantum dots; rGO—reduced graphene oxide; RhB—rhodamine B; TC—tetracycline.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Amaral, L.O.; Daniel-da-Silva, A.L. MoS2 and MoS2 Nanocomposites for Adsorption and Photodegradation of Water Pollutants: A Review. Molecules 2022, 27, 6782. https://doi.org/10.3390/molecules27206782

AMA Style

Amaral LO, Daniel-da-Silva AL. MoS2 and MoS2 Nanocomposites for Adsorption and Photodegradation of Water Pollutants: A Review. Molecules. 2022; 27(20):6782. https://doi.org/10.3390/molecules27206782

Chicago/Turabian Style

Amaral, Leonardo O., and Ana L. Daniel-da-Silva. 2022. "MoS2 and MoS2 Nanocomposites for Adsorption and Photodegradation of Water Pollutants: A Review" Molecules 27, no. 20: 6782. https://doi.org/10.3390/molecules27206782

Article Metrics

Back to TopTop